Anda di halaman 1dari 98

1

THE SCIENTIFIC MODEL OF


CAUSALITY
James J. Heckman*
Causality is a very intuitive notion that is difficult to make precise
without lapsing into tautology. Two ingredients are central to any
definition: (1) a set of possible outcomes (counterfactuals) generated
by a function of a set of factors or determinants and (2) a
manipulation where one (or more) of the factors or determinants
is changed. An effect is realized as a change in the argument of a
stable function that produces the same change in the outcome for a
class of interventions that change the factors by the same amount.
The outcomes are compared at different levels of the factors or
generating variables. Holding all factors save one at a constant
level, the change in the outcome associated with manipulation of the
varied factor is called a causal effect of the manipulated factor. This
definition, or some version of it, goes back to Mill (1848) and
Marshall (1890). Haavelmos (1943) made it more precise within the
context of linear equations models. The phrase ceteris paribus
(everything else held constant) is a mainstay of economic analysis
This research was supported by NSF 97-09-873, 00-99195, NSF SES0241858, NIH R01-HD043411, and the American Bar Foundation. An earlier
version of this paper was presented at the ISI meeting in Seoul, Korea, in August
2001. I am grateful to Jaap Abbring and Edward Vytlacil for very helpful
discussions about the topics of this paper over the past five years. Yu Xie and
especially T. N. Srinivasan made helpful comments on this version. Some of the
material in this paper also appears in Heckman and Vytlacil (2006a,b).
*University of Chicago, University College London, and the American
Bar Foundation

HECKMAN

and captures the essential idea underlying causal models. This paper
develops the scientific model of causality developed in economics and
compares it to methods advocated in epidemiology, statistics, and in
many of the social sciences outside of economics that have been
influenced by statistics and epidemiology.
I make two main points that are firmly anchored in the econometric tradition. The first is that causality is a property of a model of
hypotheticals. A fully articulated model of the phenomena being
studied precisely defines hypothetical or counterfactual states.1 A
definition of causality drops out of a fully articulated model as an
automatic by-product. A model is a set of possible counterfactual
worlds constructed under some rules. The rules may be the laws of
physics, the consequences of utility maximization, or the rules governing social interactions, to take only three of many possible examples.
A model is in the mind. As a consequence, causality is in the mind.
In order to be precise, counterfactual statements must be made
within a precisely stated model. Ambiguity in model specification
implies ambiguity in the definition of counterfactuals and hence of
the notion of causality. The more complete the model of counterfactuals, the more precise the definition of causality. The ambiguity
and controversy surrounding discussions of causal models are consequences of analysts wanting something for nothing: a definition of
causality without a clearly articulated model of the phenomenon
being described (i.e., a model of counterfactuals). They want to
describe a phenomenon as being modeled causally without producing a clear model of how the phenomenon being described is generated or what mechanisms select the counterfactuals that are observed
in hypothetical or real samples. In the words of Holland (1986), they
want to model the effects of causes without modeling the causes of
effects. Science is all about constructing models of the causes of
effects. This paper develops the scientific model of causality and
shows its value in analyzing policy problems.
My second main point is that the existing literature on causal
inference in statistics confuses three distinct tasks that need to be
carefully distinguished:
1
I will use the term counterfactual as defined in philosophy. A counterfactual need not be contrary to certain facts. It is just a hypothetical. The term
hypothetical would be better and I will use the two concepts interchangeably.

THE SCIENTIFIC MODEL OF CAUSALITY


*

Definitions of counterfactuals.
Identification of causal models from population distributions
(infinite samples without any sampling variation). The hypothetical populations producing these distributions may be subject to
selection bias, attrition, and the like. However, issues of sampling
variability of empirical distributions are irrelevant for the analysis
of this problem.
Identification of causal models from actual data, where sampling
variability is an issue. This analysis recognizes the difference
between empirical distributions based on sampled data and population distributions generating the data.

Table 1 represents these three tasks.


The first task is a matter of science, logic, and imagination. It is
also partly a matter of convention. A model of counterfactuals is
more widely accepted the more widely accepted are its ingredients,
which are
*

the rules of the derivation of a model including whether or not the


rules of logic and mathematics are followed;
its agreement with other theories; and
its agreement with the accepted interpretations of facts.

Models are not empirical statements or descriptions of actual worlds.


They are descriptions of hypothetical worlds obtained by varying
hypotheticallythe factors determining outcomes.

TABLE 1
Three Distinct Tasks Arising from Analysis of Causal Models
Task

Description

Requirements

Defining the Set of Hypotheticals


or Counterfactuals

A Scientific Theory

Identifying Parameters
(Causal or Otherwise)
from Hypothetical Population Data

Mathematical Analysis of
Point or Set Identification

Identifying Parameters from Real Data

Estimation and Testing Theory

HECKMAN

The second task is one of inference in very large samples. Can


we recover counterfactuals (or means or distributions of counterfactuals) from data that are free of sampling variation? This is the
identification problem. It abstracts from any variability in estimates
due to sampling variation. It is strictly an issue of finding unique
mappings from population distributions, population moments or
other population measures to causal parameters.
The third task is one of inference in practice. Can one recover a
given model or the desired causal parameters from a given set of data?
This entails issues of inference and testing in real world samples. This
is the task most familiar to statisticians and empirical social scientists.
This essay focuses on the first two tasks. Identification is discussed,
but issues of sampling distributions of estimators, such as efficiency,
are not.
Some of the controversy surrounding counterfactuals and causal models is partly a consequence of analysts being unclear about
these three distinct tasks and often confusing solutions to each of
them. Some analysts associate particular methods of estimation (e.g.,
matching or instrumental variable estimation) with causal inference
and the definition of causal parameters. Such associations confuse the
three distinct tasks of definition, identification, and estimation. Each
method for estimating causal parameters makes some assumptions
and forces certain constraints on the counterfactuals.
Many statisticians are uncomfortable with counterfactuals.
Their discomfort arises in part from the need to specify models to
interpret and identify counterfactuals. Most statisticians are not
trained in science or social science and adopt as their credo that
they should stick to the facts. An extreme recent example of this
discomfort is expressed by Dawid (2000), who denies the need for, or
validity of, counterfactual analysis. Tukey (1986) rejects the provisional nature of causal knowledgei.e., its dependence on a priori
models to define the universe of counterfactuals and the mechanisms
of selection and the dependence of estimators of causal parameters on
a priori, untestable assumptions.2 Cox (1992) appears to accept the
provisional nature of causal knowledge (see also Cox and Wermuth
1996). Science is based on counterfactuals and theoretical models.
2

The exchange between Heckman and Tukey in Wainer (1986) anticipates many of the issues raised in this paper.

THE SCIENTIFIC MODEL OF CAUSALITY

Human knowledge is produced by constructing counterfactuals and


theories. Blind empiricism unguided by a theoretical framework for
interpreting facts leads nowhere.
Causal models which are widely used in epidemiology and statistics are incompletely specified because they do not delineate selection
mechanisms for how hypothetical counterfactuals are realized or how
hypothetical interventions are implemented even in hypothetical
populations. They focus only on outcomes of treatment, leaving the
model-selecting outcomes only implicitly specified. In addition, in this
literature the construction of counterfactual outcomes is based on intuition and not on explicit formal models. Instead of modeling outcome
selection mechanisms, a metaphor of random selection is adopted.
This emphasis on randomization or its surrogates (like matching) rules
out a variety of alternative channels of identification of counterfactuals
from population or sample data. It has practical consequences because
of the conflation of step one with steps two and three in Table 1. Since
randomization is used to define the parameters of interest, this practice
sometimes leads to the confusion that randomization is the only way
or at least the best wayto identify causal parameters from real data. In
truth, this is not always so, as I show in this paper.
Another reason why epidemiological and statistical models are
incomplete is that they do not specify the sources of randomness generating the unobservables in the modelsi.e., they do not explain why
observationally identical people make different choices and have different outcomes given the same choice. Modeling these unobservables
greatly facilitates the choice of estimators to identify causal parameters.
Statistical and epidemiological models are incomplete because they are
recursive. They do not allow for simultaneous choices of outcomes of
treatment that are at the heart of game theory and models of social
interactions (e.g., see Tamer 2003; Brock and Durlauf 2001). They rule
out the possibility that one outcome can cause another if all outcomes
are chosen simultaneously. They are also incomplete because the ingredients of the treatments are not considered at a finer level.
Treatment is usually a black box of many aggregate factors that are
not isolated or related to underlying theory in a precise way. This makes
it difficult to understand what factor or set of factors produces the
effect of the intervention being analyzed. The treatment effects identified in the statistical literature cannot be used to forecast out-of-sample
to new populations. They are incomplete because they do not

HECKMAN

distinguish uncertainty from the point of view of the agent being


analyzed from variability as analyzed by the observing social scientist.
Economists since the time of Haavelmo (1943, 1944) have
recognized the need for precise models to construct counterfactuals
and to answer causal questions and more general policy evaluation
questions, including making out-of-sample forecasts. The econometric
framework is explicit about how counterfactuals are generated and
how interventions are assigned (the rules of assigning treatment).
The sources of unobservables, in both treatment assignment equations and outcome equations, and the relationship between the unobservables are studied. Rather than leaving the rule governing selection
of treatment implicit, the econometric approach explicitly models the
relationship between the unobservables in outcome equations and
selection equations to identify causal models from data and to clarify
the nature of identifying assumptions. The theory of structural modeling in econometrics is based on these principles.
The goal of the econometric literature, like the goal of all
science, is to model phenomena at a deeper level, to understand the
causes producing the effects so that we can use empirical versions of
the models to forecast the effects of interventions never previously
experienced, to calculate a variety of policy counterfactuals, and to
use scientific theory to guide the choices of estimators and the interpretation of the evidence. These activities require development of a
more elaborate theory than is envisioned in the current literature on
causal inference in epidemiology and statistics.
This essay is in five parts. Section 1 discusses policy evaluation
questions as a backdrop against which to compare alternative
approaches to causal inference. A notation is developed and both individual-level and population-level causal effects are defined. Populationlevel effects are defined both in terms of means and distributions.
Uncertainty at the individual level is introduced to account for one
source of randomness across persons in terms of outcomes and choices.
Section 2 is the heart of the paper. It defines causality using
structural econometric models and analyzes both objective outcomes
and subjective evaluations. It defines structural models and policyinvariant structural parameters. A definition of causality in models
with simultaneously determined outcomes is presented. A distinction
between conditioning and fixing variables is developed. The Neyman
(1923)Rubin (1978) model advocated in statistics is compared to the

THE SCIENTIFIC MODEL OF CAUSALITY

scientific model. Marschaks maxim is defined. This maxim links the


statistical treatment effect literature to the literature on structural
models by showing that statistical treatment effects focus on answering
one narrow question while the structural approach attempts to answer
many questions. It is usually easier to answer one question well than to
answer many questions at the same time but the narrowness of the
question answered in the treatment effect literature limits the applicability of the answer obtained to address other questions.
Section 3 briefly discusses the identification problem at a general level (task 2 in Table 1). Section 4 applies the framework of the
paper to the identification of four widely used estimators for causal
inference and the implicit identifying assumptions that justify their
application. This section is only intended as a comprehensive survey.
Section 5 concludes.

1. POLICY EVALUATION QUESTIONS AND CRITERIA OF


INTEREST
This paper discusses questions of causal inference in terms of
policy evaluation and policy forecasting problems. Such a focus
appears to limit the scope of the inquiry. In fact, it makes the discussion more precise by placing it in a concrete context. By focusing on
policy questions, the discussion gains tangibility, something often
lacking in the literature on causality. In social science, a major use of
causal analysis is in determining effects of various policies. Causal
analysis is almost always directed toward answering policy questions.
This section first presents three central policy evaluation questions. It then defines the notation used in this paper and the definition
of individual-level causal effects or treatment effects. The policy evaluation problem is discussed in general terms. Population-level mean
treatment parameters are then defined and distributional criteria are
also presented. We discuss, in general terms, the type of data needed
to construct the policy evaluation criteria.
1.1. Three Policy Evaluation Problems
Three broad classes of policy evaluation questions are of general
interest. Policy evaluation question one is:

HECKMAN

P1: Evaluating the impact of historical interventions on outcomes including their impact in terms of welfare.
By historical, I refer to interventions actually experienced. A variety of
outcomes and welfare criteria might be used to form these evaluations.
By impact, I mean constructing either individual-level or populationlevel counterfactuals and their valuations. By welfare, I mean the valuations of the outcomes obtained from the intervention by the agents being
analyzed or some other party (e.g., the parents of the agent).
P1 is the problem of internal validity. It is the problem of
identifying a given treatment parameter or a set of treatment parameters in a given environment (see Campbell and Stanley 1963). This
is the policy question addressed in the epidemiological and statistical
literature on causality. A drug trial for a particular patient population
is the prototypical problem in that literature.
Most policy evaluation is designed with an eye toward the future
and toward decisions about new policies and application of old policies
to new environments. I distinguish a second task of policy analysis:
P2: Forecasting the impacts (constructing counterfactual
states) of interventions implemented in one environment in other environments, including their impacts in terms of welfare.
Included in these interventions are policies described by generic characteristics (e.g., tax or benefit rates, etc.) that are applied to different
groups of people or in different time periods from those studied in
previous implementations of these policies. This is the problem of
external validity: taking a treatment parameter or a set of parameters
estimated in one environment to another environment. The environment includes the characteristics of individuals and of their social
and economic setting.
Finally, the most ambitious problem is forecasting the effect of
a new policy, never previously experienced:
P3: Forecasting the impacts of interventions (constructing counterfactual states associated with interventions) never historically experienced to other environments, including their impacts in terms of welfare.
This problem requires that one use past history to forecast the
consequences of new policies. It is a fundamental problem in

THE SCIENTIFIC MODEL OF CAUSALITY

knowledge.3 I now present a framework within which one can address


these problems in a systematic fashion. It is also a framework that can
be used for causal inference.
1.2. Notation and Definition of Individual-Level Treatment or Causal
Effects
To evaluate is to value and to compare values among possible outcomes.
These are two distinct tasks that I distinguish in this essay. Define outcomes corresponding to state (policy, treatment) s for person ! as Y(s, !),
! 2 . One can think of  as a universe of individuals each characterized
by their own element !. The ! encompass all features of individuals that
affect Y outcomes. Y(s, !) may be generated from a scientific or social
science theory. Y(s, !) may be vector valued. The components of Y(s, !)
may also be interdependent, as in the Cowles Commission simultaneous
equations model developed by Haavelmo (1943, 1944) and discussed in
Section 2. The components of Y(s, !) may be discrete, continuous, or
mixed discrete-continuous random variables.
I use ! as a shorthand descriptor of the state of a person. We
(the analyst) may observe variables X(!) that characterize the person
as well. In addition, there may be model unobservables. I develop this
distinction further in Section 2.
The Y(s, !) are outcomes after treatment is chosen. In advance
of treatment, agents may not know the Y(s, !) but may make forecasts
about them. These forecasts may influence their decisions to participate
in a treatment or may influence the agents who make decisions about
whether or not an individual participates in the treatment. Selection
into the program based on actual or anticipated components of outcomes gives rise to the selection problem in the evaluation literature.
Let S be the set of possible treatments denoted by s. For
simplicity of exposition, I assume that this set is the same for all !.4
For each choice of s 2 S and for each person !, we obtain a collection
of possible outcomes given by fY(s, !)gs2S . The set S may be finite
3

Knight (1921:313) succinctly summarizes the problem: The existence


of a problem in knowledge depends on the future being different from the past,
while the possibility of a solution of the problem depends on the future being like
the past.
4
At the cost of a more cumbersome notation, this assumption can be
modified so that S sets are !-specific.

10

HECKMAN

(e.g., J states with S {1, . . . J}), countable, or may be defined on


the continuum (e.g., S 0, 1) so there are an uncountable number
of states. For example, if S f0, 1g, there are two policies (or
treatments), one of which may be a no-treatment statefor example,
Y(0, !) is the outcome for a person ! not getting a treatment like a
drug, schooling, or access to a new technology, while Y(1, !) corresponds to person ! getting the drug, schooling or access.
Each state (treatment, policy) may consist of a compound of
subcomponent states. In this case, we can define s as a vector (e.g.,
s (s1, s2, . . . , sk)) corresponding to the different components that
comprise treatment. Thus a job training program typically consists of
a package of treatments. We might be interested in the package or one
(or more) of its components. Thus s1 may be months of vocational
education, s2 quality of training and so forth. The outcomes may be
time subscripted as well, with Yt (s, !) corresponding to outcomes of
treatment measured at different times. The index set for t may be the
integers, corresponding to discrete time, or an interval, corresponding to
continuous time.5 The Yt (s, !) are realized or ex post (after treatment)
outcomes. When choosing treatment, these values may not be known.
Gill and Robins (2001), Abbring and Van Den Berg (2003), Lechner
(2004), Heckman and Vytlacil (2006a,b), and Heckman and Navarro
(2006) develop models for dynamic counterfactuals.
Each policy regime p 2 P consists of a collection of possible
treatments S p  S. Different policy regimes may include some of the
same subsets of S. Associated with each policy regime is an assignment
mechanism  2 T p , where T p is the set of possible mechanisms under
policy p. (Some policy regimes may rule out some assignment mechanisms.) The assignment mechanism determines the allocation of persons
! 2  to treatment. It implicitly sets the scale of the program. The
mechanism could include randomization so that the assignment mechanism would assign probabilities s 2 0; 1 to each treatment s 2 S p . Let
p denote the set of families s s 2 Sp , s 2 [0, 1], such that s 2 Sp s 1.
Then,
p : T p ! p ;
5
In principle, in addition to indexing S by ! (S ! ) so there are personspecific treatment possibility sets, we could index by t (S !;t ), but we assume, for
simplicity, a common S for all ! and t.

11

THE SCIENTIFIC MODEL OF CAUSALITY

where p(!, ) 2 


 p is a family of probabilities which we note
alternatively s ! s 2 Sp . This signifies that, under policy p with
assignment mechanism , person ! receives treatment sp with probability sp !. For each person !, the special case of deterministic
assignment sets s0 ! 1 for exactly one treatment s0 2 S p and sets
s0 ! 0 for all s 2 S p nfs0 g.
For deterministic policy assignment rules, a universal policy
may consist of a single treatment (S p may consist of a single element).
Treatment can include direct receipt of some intervention (e.g., a
drug, education) as well as the tax payment for financing the treatment. For some persons, the assigned treatment may only be the tax
payment. In the special case where some get no treatment (! 2 0)
and others get treatment (! 2 1), and there are two elements in S p
(e.g., S p f0; 1g), we produce the classical binary treatment-control
comparison.
Two assumptions are often invoked in the literature.6 In our
notation, they are:
Ys; !; p;  Ys; !; p0 ;  Ys; !;  for s 2 S p \ S p0 ;
 2 T p \ T p0 ; for all p; p0 2 P and ! 2 

A-1

This assumption says that outcomes for person ! under treatment s with
assignment mechanism  are the same in two different policy regimes
which both include s as a possible treatment. It rules out social interactions and general equilibrium effects. A second assumption rules out any
effect of the assignment mechanism on potential outcomes.
Irrespective of assignment mechanism ; for all policies
p 2 P; Ys; !;  Ys; ! for all s 2 S p and
! 2 ; so the outcome is not affected by the assignment:

A-2

This assumption maintains that the outcome is the same no matter


what the choice of assignment mechanism. (A-2) rules out, among
other things, the phenomenon of randomization bias discussed in
Heckman, LaLonde, and Smith (1999) where agent behavior is

See, e.g., Holland (1986) or Rubin (1986).

12

HECKMAN

affected by the act of participating in an experiment. Such effects are


also called Hawthorne effects.
Heckman, LaLonde, and Smith (1999) discuss the evidence
against both assumptions. In much of this essay, I maintain these
strong assumptions mostly to simplify the discussion. But the reader
should be aware of the strong limitations imposed by these assumptions. Recent work in economics tests and relaxes these assumptions
(see Heckman and Vytlacil 2006a).
Under these assumptions, the individual-level treatment effect
for person ! comparing outcomes from treatment s with outcomes
from treatment s0 is
Ys; !  Ys0 ; !;

s 6 s0 ;

where two elements are selected s, s0 2 S.7 This is also called an


individual-level causal effect. This may be a random variable or a
constant. Our framework accommodates both interpretations. Thus
the same individual with the same choice set and characteristics may
have the same outcome in a sequence of trials or it may be random
across trials. We discuss intrinsic variability at the individual level in
Section 2.8
Other comparisons might be made. Comparisons can be made
in terms of utilities (personal, V(Y(s, !), !), or in terms of planner
preferences, VG). Thus one can ask if V(Y(s, !), !) > V(Y(s0 , !), !)
or not (is the person better off as a result of treatment s compared
to treatment s0 ?) Treatments s and s0 may be bundles of components

One could define the treatment effect more generally as


Ys; !; p;   Ys0 ; !; p; :

This makes clear that the policy treatment effect is defined under a particular
policy regime and for a particular mechanism of selection within a policy regime.
One could define treatment effects for policy regimes or regime selection mechanisms by varying the arguments p or  respectively, holding the other arguments
fixed.
8
There is a disagreement in the literature on whether or not the individual-level treatment effects are constants or random at the individual level. I
develop both cases in this paper.

13

THE SCIENTIFIC MODEL OF CAUSALITY

as previously discussed. One could define the treatment effect as


1[V(Y (s, !), !) > V(Y(s0 , !), !)] where 1[] 1 if the argument in
brackets is true and is zero otherwise. These definitions of treatment
effects embody Marshalls notion of ceteris paribus. Holding !
fixed holds all features about the person fixed except the treatment
assigned s.9
Social welfare theory constructs aggregates over  or subsets
of  (Sen 1999). A comparison of two policies {sp(!)}!2 and
{sp0 (!)}!2, using the social welfare function VG({Y(s(!), !)}!2),
can be expressed as
VG fYsp ; !; !g!2  VG fYsp0 ; !; !g!2 :
We can use an indicator function to denote when this term is positive:
1[VG({Y(sp(!), !)}!2) > VG({Y(sp0 (!), !)}!2)]. A special case of this
analysis is cost-benefit analysis in economics where willingness to pay
measures W(s(!), !) are associated with each person. The cost-benefit
comparison of two policies is
Cost Benefit : CBp;p0

WYsp !; !d!

WYsp0 !; !d!;

One might compare outcomes in different sets that are ordered. Thus,
for a particular policy regime and assignment mechanism, if Y (s, !) is scalar
income and we compare outcomes for s 2 S A with outcomes for s0 2 S B , where
S A \ S B , then one might compare YsA  YsB , where
sA arg maxYs; !
s 2 SA

and

sB arg maxYs; !:
s 2 SB

This compares the best in one choice set with the best in the other. A particular
case is the comparison of the best choice with the next best choice. To do so,
define s0 arg maxs2S Ys; !; S B S n fs0 g and define the treatment effect as
Ys0  YsB . This is the comparison of the highest outcome over S with the next best
outcome. In principle, many different individual level comparisons might be
constructed, and they may be computed using personal preferences, V!, using
the preferences of the planner, VG, or using the preferences of the planner over
preferences of agents.

14

HECKMAN

where p, p0 are two different policies, p0 may correspond to a benchmark


of no policy, and  (!) is the distribution of !.10 The distribution  (!)
is constructed over the individual characteristics ! (e.g., age, sex,
race, income) The Benthamite criterion replaces W(Y(s(!), !)) with
V(Y(s(!), !)) in the preceding expressions and integrates utilities across
persons:
Benthamite : Bp;p0

VYsp !; !d! 


VYsp0 !; !d!:


I now discuss a fundamental problem that arises in constructing these and other criteria from data. This takes me to the problem of
causal inference, the second task delineated in Table 1. Recall that
I am talking about inference in a population, not in a sample, so no
issues of sampling variability arise.
1.3. The Evaluation Problem
Operating purely in the domain of theory, I have assumed a world with a
well-defined set of individuals ! 2  and a universe of counterfactuals or
hypotheticals defined for each person Y (s, !), s 2 S. Different policies
p 2 P select treatment for persons. Each policy can in principle
assign treatment to persons by different mechanisms  2 T . In the
absence of a theory, there are no well-defined rules for constructing
counterfactual or hypothetical states or constructing the assignment to
treatment rules p .11 Scientific theories provide algorithms for
generating the universe of internally consistent, theory-consistent counterfactual states.
These hypothetical states are possible worlds. They are products of a purely mental activity. No empirical problem arises in
constructing these theoretically possible worlds. Indeed, in forecasting
new policies, or projecting the effects of old policies to new
10
These willingness-to-pay measures are standard in the economics
literature (e.g., see Boadway and Bruce 1984).
11
Efforts like those of Lewis (1974) to define admissible counterfactual
states without an articulated theory as closest possible worlds founder on the
lack of any meaningful metric or topology to measure closeness among possible
worlds. Statisticians often appeal to this theory, but it is not operational (e.g., see
Gill and Robins 2001 for one such appeal).

THE SCIENTIFIC MODEL OF CAUSALITY

15

environments, some of the Y(s, !) may have never been observed for
anyone. Different theories produce different outcomes Y(s, !) and
different p (!).
The evaluation problem, in contrast to the model construction
problem, is an identification problem that arises in constructing the
counterfactual states and treatment assignment rules produced by
abstract models from population data. This is the second task presented in Table 1.
This problem is not precisely stated until the data available to the
analyst are precisely defined. Different subfields in science and social
science assume access to different types of data. They also make different
assumptions about the underlying models generating the counterfactuals
and mechanisms for selecting which counterfactuals are actually
observed.
At any point in time, we can observe person ! in one state but
not in any of the other states. The states are mutually exclusive. Thus
we do not observe Y(s0 , !) for person ! if we observe Y(s, !), s 6 s0 .
Let D(s, !) 1 if we observe person ! in state s. Then D(s0 , !) 0
for s 6 s0 . D(s, !) is generated by p (!) : D(s, !) 1 if p (!) s.
We observe Y(s, !), if D(s, !) 1 but we do not observe
Y(s0 , !), s 6 s0 . We can define observed Y(!) as
Y!

Ds; !Ys; !:12

s2S

Without further assumptions, constructing an empirical counterpart to


equation (1) is impossible from the data on (Y(!), D(!)), ! 2 . This
formulation of the evaluation problem is known as Quandts switching
regression model (Quandt 1958, 1974) and is attributed in statistics to
Neyman (1923), Cox (1958), and Rubin (1978). A revision of it is
formulated in a linear equations context for a continuum of treatments
by Haavelmo (1943). The Roy model (Roy 1951) in economics is
another version of it with two possible treatment outcomes
(S f0,1g) and a scalar outcome measure and a particular selection
mechanism  2 T which is that D(1, !) 1 (Y(1, !) > Y(0, !)) where
1[] is an indicator function which equals 1 when the event inside the
12

In the general case, Y(!)


Dirac function.

D(s, !)Y(s, !)ds where D (s, !) is a

16

HECKMAN

parentheses is true and is zero otherwise.13 The mechanism of selection


depends on the potential outcomes. Agents choose the sector with
the highest outcome, so the actual selection mechanism is not a
randomization.
Social experiments attempt to create assignment rules so that
D(s, !) is random with respect to fY(s, !)gs2S for each ! (i.e., so that
receipt of treatment is independent of the outcome of treatment). When
agents self-select into treatment, rather than being randomly assigned, in
general the D(s, !) are not independent of fY(s, !)gs2S . This arises in
the Roy model example. This selection rule creates the potential for selfselection bias in inference. We discuss this problem at length in Section 4.
The problem of self-selection is an essential aspect of the evaluation
problem when data are generated by choices of agents. The agents making
choices may be different from the agents receiving treatment (e.g., parents
making choices for children). Such choices can include compliance with the
protocols of a social experiment as well as ordinary choices about outcomes
that people make in everyday life. Observe that in the Roy model, the
choice of treatment (including the decision not to attrite from a program) is
informative on the relative valuation of the Y(s, !). This point is more
general and receives considerable emphasis in the econometric literature
but none in the statistical or epidemiological literature. Choices of treatment provide information on subjective relative evaluations of treatment
by the decision maker and provides analysts with information on agent
valuations of outcomes that are of independent interest.
A central problem considered in the literature on causal inference is the absence of information on outcomes for person ! other
than the outcome that is observed. Even a perfectly implemented
social experiment does not solve this problem (Heckman 1992) and,
even under ideal conditions, randomization identifies only one component of fY(s, !)gs2S . In addition, even with ideal data and infinite
samples some of the s 2 S may not be observed if one is seeking to
evaluate policies that produce new outcome states.
There are two main avenues of escape from this problem. The
first, featured in explicitly formulated econometric models, often
called structural econometric analysis, is to model Y(s, !) explicitly
in terms of its determinants as specified by theory. This entails
13

In terms of the assignment mechanism, p (!, ) 1 for ! such that


Y(1, !) > Y(0, !).

THE SCIENTIFIC MODEL OF CAUSALITY

17

describing ! and carefully distinguishing what agents know and what


the analyst knows. This approach also models D(s, !)or p (!)
and the dependence between Y(s, !) and D(s, !) produced from
variables common to Y(s, !) and D(s, !). The Roy model, previously
discussed, explicitly models this dependence.14 Like all scientific
models, this approach seeks to understand the factors underlying
outcome, choice of outcome equations, and their relationship.
Empirical models explicitly based on economic or social theory pursue this avenue of investigation. Some statisticians call this the scientific approach and are surprisingly hostile to it (Holland 1986).15
A second avenue of escape, and the one pursued in the recent
epidemiological and statistical treatment effect literature, defines the
problem away from estimating Y(s, !) to be one of estimating some
population version of equation (1), most often a mean, without modeling
those factors giving rise to the outcome or the relationship between the
outcomes and the mechanism selecting outcomes. Agent valuations of
outcomes are ignored. The treatment effect literature focuses almost
exclusively on policy problem P1 for the subset of outcomes that is
observed. It ignores the problems of forecasting a policy in a new environment (problem P2) or a policy never previously experienced (problem
P3). Forecasting the effects of new policies is a central task of science and
public policy analysis that the treatment effect literature ignores.16
1.4. Population-Level Treatment Parameters
Constructing equation (1) or any of the other individual-level parameters
defined in Section 1.2 for a given person is a difficult task because we rarely
observe the same person ! in distinct s states. In addition, some of the states
in S may not be experienced by anyone. The conventional approach in
the treatment effect literature is to reformulate the parameter of interest
to be some summary measure of the population distribution of treatment
14

See Heckman and Honore (1990) for a discussion of this model.


I include in this approach methods based on panel data or more
generally the method of paired comparisons, as applications of the scientific
approaches. Under special conditions discussed in Heckman and Smith (1998),
we can observe the same person in states s and s0 in different time periods and can
construct (1) for all !.
16
See Heckman and Vytlacil (2005) for one synthesis of the treatment
effect and the structural literatures.
15

18

HECKMAN

effects, most often a mean, or sometimes the distribution itself, rather


than attempting to identify individual treatment effects. This approach
focuses on presenting some summary measure of outcomes, not analyzing determinants of outcomes.17 This approach also confines attention to
the subsets of S that are observed states. Thus the objects of interest are
redefined to be the distributions of (Y( j, !)  Y(k, !)) over !, conditional on known components of !, or certain means (or quantiles) of the
distribution of (Y( j, !)  Y(k, !)) over !, conditional on known components of ! (Heckman, Smith, and Clements 1997) or of Y( j, !) and
Y(k, !) separately (Abadie, Angrist, and Imbens 2002). The standard
assumptions in the treatment effect literature are that all states in S are
observed, and that assumptions (A-1) and (A-2) hold (see Holland 1986;
Rubin 1986).
The conventional parameter of interest, and the focus of many
investigations in economics and statistics, is the average treatment
effect (ATE). For program (treatment) j compared to program
(treatment) k, this parameter is
ATE j; k E! Y j; !  Yk; !;

3a

where E! means that we take expectations with respect to distribution of


the factors generating outcomes and choices that characterize !.
Conditioning on covariates X, which are observed components associated
with ! (and hence working with conditional distributions), this parameter is
ATE j; k j x E! Y j; !  Yk; ! j X x:

3b

This is the effect of assigning a person to a treatmenttaking someone


from the overall population (3a) or a subpopulation conditional on X (3b)
and determining the mean gain of the move from base state k, averaging
over the factors that determine Y but are not captured by X. This
parameter is also the effect of moving the society from a universal policy
(characterized by policy k) and moving to a universal policy of j (e.g.,
from no social security to full population coverage). Such a policy would
likely induce social interaction and general equilibrium effects that are

17

The effects of causes and not the causes of effects, in the language of
Holland (1986).

THE SCIENTIFIC MODEL OF CAUSALITY

19

assumed away by (A-1) in the treatment effect literature and which, if


present, fundamentally alter the interpretation placed on this parameter.
A second conventional parameter in this literature is the average effect of treatment on the treated. Letting D( j, !) 1 denote
receipt of treatment j, the conventional parameter is
TT j; k E! Y j; !  Yk; ! j D j; ! 1:

4a

For a population conditional on X x, it is


TT j; k j x E! Y j; !  Yk; ! j D j; ! 1; X x:

4b

These are, respectively, the mean impact of moving persons from k to


j for those people who get treatment, unconditional and conditional
on X x.
A parallel pair of parameters for nonparticipants is treatment
on the untreated, where D( j, !) 0 denotes no treatment at level j:
TUT j; k E! Y j; !  Yk; ! j D j; ! 0
TUT j; k j x E! Y j; !  Yk; ! j D j; ! 0; X x:

5a
5b

These parameters answer (conditionally and unconditionally) the


question of how extension of a program to nonparticipants as a
group would affect their outcomes.18
The population treatment parameters just discussed are average effects: how the average in one treatment group compares with
the average for another. The distinction between the marginal and
average return has wide applicability in many areas of social science.
The average student going to college may have higher earnings than
the marginal student who is indifferent between going to school or
not. It is often of interest to evaluate the impact of marginal extensions (or contractions) of a program. Incremental cost-benefit analysis is conducted in terms of marginal gains and benefits. The effect of
treatment for people at the margin of indifference (EOTM) between

18

Analogous to the pairwise comparisons, we can define setwise


comparisons as is done in footnote 9.

20

HECKMAN

j and k, given that these are the best two choices available is, with
respect to personal preferences, and with respect to choice-specific
costs P ( j, !),
EOTMV
! Y j; !  Yk; !

0
1
 VY j; !; P j; !; ! VYk; !; Pk; !; !;

9

B
C
 VY j; !; P j; !; ! =
B
C

B
C:

B
 VYl; !; Pl; !; !; C
E! BY j; !  Yk; !
C
;
 VY k; !; P k; !; !
B
C

@
A


l 6 j; k

6
This is the mean gain to people indifferent between j and k, given that
these are the best two options available. In a parallel fashion, we can
G
define EOTMV
! (Y( j)  Y(k)) using the preferences of another person
(e.g., the parent of a child or a paternalistic bureaucrat).19
A generalization of this parameter called the marginal treatment effectdeveloped in Heckman and Vytlacil (1999, 2000, 2005,
2006b), Heckman (2001), and estimated in Carneiro, Heckman, and
Vytlacil (2005)plays a central role in organizing and interpreting a
wide variety of evaluation estimators. Many other mean treatment
parameters can be defined depending on the choice of the conditioning set. Analogous definitions can be given for median and other
quantile versions of these parameters (see Heckman, Smith, and
Clements 1997; Abadie, Angrist, and Imbens 2002). Although means
are conventional, distributions of treatment parameters are also of
considerable interest, and we consider them in the next section.
Mean treatment effects play a special role in the statistical
approach to causality. They are the centerpiece of the Rubin (1986)
Holland (1986) model and in many other studies in statistics and
epidemiology. Social experiments with full compliance and no disruption can identify these means because of a special mathematical
property of means. If we can identify the mean of Y( j, !) and the
mean of Y(k, !) from an experiment where j is the treatment and k is
the baseline, we can form the average treatment effect for j compared
19

An analogous parameter can be defined for mean setwise comparisons


as in footnote 9.

THE SCIENTIFIC MODEL OF CAUSALITY

21

with k (3a). These can be formed over two different groups of people
classified by their X values. By a similar argument, we can form the
treatment on the treated parameter (TT) (4a) or (TUT) (5a) by
randomizing over particular subsets of the population (D 1 or
D 0, respectively) assuming full compliance and no randomization
(disruption) bias. Disruption bias arises when the experiment itself
affects outcomes (Y(s, !))!2 and (A-2) is violated.20
The case for randomization is weaker if the analyst is interested
in other summary measures of the distribution, or the distribution
itself. Experiments do not solve the problem that we cannot form
Y (s, !)  Y(s0 , !) for any person. Randomization is not an effective
procedure for identifying median gains, or the distribution of gains,
under general conditions. The elevation of population means to be the
central population-level causal parameters promotes randomization
as an ideal estimation method. By focusing exclusively on mean outcomes, the statistical literature converts a metaphor for outcome
selectionrandomizationinto an ideal.
1.5. Criteria of Interest Besides the Mean: Distributions of
Counterfactuals
Although means are traditional, the answer to many interesting policy
evaluation questions requires knowledge of features of the distribution of program gains other than some mean. It is also of interest to
know the following for scalar outcomes
a.

The proportion of people taking the program j who benefit from it


relative to some alternative k, Pr!(Y( j, !) > Y(k, !)jD( j, !) 1);
b. The proportion of the total population that benefits from the
program k compared with program j, Pr!(Y ( j, !) > Y (k, !)),
sometimes called the voting criterion;
c. Selected quantiles of the impact distribution;21
d. The distribution of gains at selected base state values, (the distribution of Y( j, !)  Y(k, !) given Y(k, !) y(k)).
20

Such disruptions leading to changed outcomes are also called


Hawthorne effects; see Heckman (1992) and Heckman, LaLonde, and Smith (1999).
21
inf { : F ()  q} where q is a quantile of the distribution and F is
the distribution function of  Y( j, !)  Y(k, !).

22

HECKMAN

Each of these measures can be defined conditional on observed characteristics X. Measure (a) is of interest in determining how widely
program gains are distributed among participants. Voters in an electorate in a democratic society are unlikely to assign the same weight
to two programs with the same mean outcome, one of which produced large favorable outcomes for only a few persons while the other
distributed smaller gains more broadly. This issue is especially relevant if program benefits are not transferrable or if restrictions on
feasible social redistributions prevent distributional objectives from
being attained.
Measure (b) is the proportion of the entire population that
benefits from a program. In a study of the political economy of
interest groups, it is useful to know which groups benefit from a
program and how widely distributed the program benefits are.
Measure (c) reveals the gains at different percentiles of the impact
distribution. Criterion (d) focuses on the distribution of impacts for
subgroups of participants with particular outcomes in the nonparticipation state. Concerns about the impact of policies on the disadvantaged emphasize such criteria (Rawls 1971). All of these measures
require knowledge of features of the joint distribution of outcomes
for participants for their construction, not just the mean. Identifying
distributions is a more demanding task than identifying means.
Distributions of counterfactuals are also required in computing
the option values conferred by social programs.22 Heckman and
Smith (1998), Aakvik, Heckman, and Vytlacil (1999, 2005),
Carneiro, Hansen, and Heckman (2001, 2003), and Cunha,
Heckman, and Navarro (2005a) develop methods for identifying
distributions of counterfactuals.
1.6. Accounting for Private and Social Uncertainty
Persons do not know the outcomes associated with possible states not
yet experienced. If some potential outcomes are not known at the time
treatment decisions are made, the best that agents can do is to forecast
them with some rule. Even if, ex post, agents know their outcome in a
benchmark state, they may not know it ex ante, and they may always
22

Heckman, Smith, and Clements (1997) present estimates of the option


values of social programs.

THE SCIENTIFIC MODEL OF CAUSALITY

23

be uncertain about what they would have experienced in alternative


states. This creates a further distinction between ex ante and ex post
evaluations of both subjective and objective outcomes. This distinction is missing from the statistical treatment effect literature.
In the literature on social choice, one form of decision-making
under uncertainty plays a central role. The Veil of Ignorance of
Vickrey (1945, 1960) and Harsanyi (1955, 1975) postulates that individuals are completely uncertain about their position in the distribution of outcomes under each policy considered, or should act as if
they are completely uncertain, and they should use expected utility
criteria (Vickrey-Harsanyi) or a maximin strategy (Rawls 1971) to
evaluate their welfare under alternative policies. Central to this viewpoint is the anonymity postulate that claims the irrelevance of any
particular persons outcome to the overall evaluation of social welfare. This form of ignorance is sometimes justified as an ethically
correct position that captures how an objectively detached observer
should evaluate alternative policies even if actual participants in the
political process use other criteria. An approach based on the Veil of
Ignorance is widely used in applied work in evaluating different
income distributions (see Foster and Sen 1998). It only requires
information about the marginal distributions of outcomes produced
under different policies. If the outcome is income, policy j is preferred
to policy k if the income distribution under j stochastically dominates
the income under k.23
An alternative criterion is required if it is desired to model
social choices where persons act in their own self-interest, or in the
interest of certain other groups (e.g., the poor, the less able) and have
at least partial knowledge about how they (or the groups they are
interested in) will fare under different policies. The outcomes in
different regimes may be dependent so that persons who benefit
under one policy may also benefit under another (see Carneiro,
Hansen, and Heckman 2001, 2003).
Because agents typically do not possess perfect information,
the simple voting criterion assuming perfect foresight discussed in
Section 1.5 may not accurately predict choices and requires
23
See Foster and Sen (1998) for a definition of stochastic dominance. It
compares one distribution with another and determines which, if either, has more
mass at favorable outcomes.

24

HECKMAN

modification. Let I ! denote the information set available to agent !.


The agent evaluates policy j against k using that information. Under
an expected utility criterion, person ! prefers policy j over k if
E! VY j; !; ! j I ! > E! VYk; !; ! j I ! :
The proportion of people who prefer j is

Z 
E! VY j; !; !jI ! >
PB j j j; k 1
d!;
E! VYk; !; ! j I !

where (!) is the distribution of ! in the population.24 The voting


criterion previously discussed in Section 1.5 is the special case where
I! (Y( j, !),Y(k, !)), so there is no uncertainty about Y( j, !) and
Y(k, !). In the more general case, the expectation is computed against
the distribution of (E! (V(Y(j, !), !jI ! )), E! (V(Y(k, !), !)jI ! )).25
Accounting for uncertainty in the analysis makes it essential to
distinguish between ex ante and ex post evaluations. Ex post, part of
the uncertainty about policy outcomes is resolved although individuals do not, in general, have full information about what their
potential outcomes would have been in policy regimes they have not
experienced and may have only incomplete information about the
policy they have experienced (e.g., the policy may have long run
consequences extending after the point of evaluation). It is useful to
index the information set I ! by t, I !,t , to recognize that information
about the outcomes of policies may accrue over time. Ex ante and
ex post assessments of a voluntary program need not agree. Ex post
assessments of a program through surveys administered to persons
who have completed it (see Katz, Gutek, Kahn, and Barton 1975)
may disagree with ex ante assessments of the program. Both may
reflect honest valuations of the program but they are reported when
agents have different information about it or have their preferences

24

Persons would not necessarily vote honestly, although in a binary


choice setting they do and there is no scope for strategic manipulation of votes
(see Moulin 1983). PB is simply a measure of relative satisfaction and need not
describe a voting outcome when other factors come into play.
25
See Cunha, Heckman, and Navarro (2005b) for computations regarding both types of joint distributions.

THE SCIENTIFIC MODEL OF CAUSALITY

25

altered by participating in the program. Before participating in a


program, persons may be uncertain about the consequences of participation. A person who has completed program j may know Y( j, !)
but can only guess at the alternative outcome Y(k, !) which they have
not experienced. In this case, ex post satisfaction with j relative to k
for agent ! is synonymous with the inequality
VY j; !; ! > E! VYk; !; ! j I ! ;

where the information is post-treatment. Survey questionnaires about


client satisfaction with a program may capture subjective elements
of program experience not captured by objective measures of outcomes that usually exclude psychic costs and benefits. (Heckman,
Smith, and Clements 1997 and Heckman and Smith 1998 present
evidence on this question.) Carneiro, Hansen, and Heckman (2001,
2003), Cunha, Heckman, and Navarro (2005a,b), and Heckman and
Navarro (2004, 2006) develop econometric methods for distinguishing
ex ante from ex post evaluations of programs.
1.7. Information Needed to Construct Various Criteria
Four ingredients are required to implement the criteria discussed in
this section: (1) private preferences, including preferences over outcomes by the decision maker; (2) social preferences, as exemplified by
social welfare function VG({Y(sp(!), !)}!2); (3) distributions of
outcomes in alternative states, and for some criteria, such as the
voting criterion, joint distributions of outcomes across policy states;
and (4) ex ante and ex post information about outcomes. Cost-benefit
analysis requires only information about means of measured outcomes and for that reason is easier to implement. The treatment effect
literature in epidemiology and statistics largely focuses on means.
Recent work in econometrics analyzes distributions of treatment
effects (see Heckman, Smith, and Clements 1997; Carneiro, Hansen,
and Heckman 2001, 2003; Cunha, Heckman, and Navarro 2005a).
The rich set of questions addressed in this section contrasts sharply
with the focus on mean outcome parameters in the epidemiology and
statistics literatures, which ignore private and social preferences and
ignore distributions of outcomes. Carneiro, Hansen, and Heckman
(2001, 2003), Cunha, Heckman, and Navarro (2005a,b), and

26

HECKMAN

Heckman and Navarro (2006) present methods for extracting private


information on evaluations and their evolution over time. I now
exposit more formally the econometric approach to formulating
causal models.

2. COUNTERFACTUALS, CAUSALITY, AND STRUCTURAL


ECONOMETRIC MODELS
This section formally defines structural models as devices for generating counterfactuals. I consider both outcome and treatment choice
equations. The scientific model of econometrics is compared with the
Neyman (1923)Rubin (1978) model of causality that dominates discussions in epidemiology, in statistics, and in certain social sciences
outside of economics. The structural equations approach and
treatment effects approach are compared and evaluated.

2.1. Generating Counterfactuals


The treatment effect and structural approaches differ in the detail
with which they specify counterfactual outcomes, Y(s, !). The scientific approach embodied in the structural economics literature models
the counterfactuals more explicitly than is common in the statistical
treatment effect literature. This facilitates the application of theory to
provide interpretation of counterfactuals and comparison of counterfactuals across empirical studies using basic parameters of social
theory. These models also suggest strategies for identifying
parameters (task 2 in Table 1). Models for counterfactuals are
the basis for extending historically experienced policies to new environments and for forecasting the effects of new policies never previously experienced. These are policy questions P2 and P3 stated in
Section 1.
Models for counterfactuals are in the mind. They are internally
consistent frameworks derived from theory. Verification and identification of these models from data are separate tasks from the purely
theoretical act of constructing internally consistent models. No issue
of sampling, inference, or selection bias is entailed in constructing
theoretical models for counterfactuals.

THE SCIENTIFIC MODEL OF CAUSALITY

27

The traditional model of econometrics is the all causes


model.26 It writes outcomes as a deterministic function of inputs:
ys gs x; us ;

where x and us are fixed variables specified by the relevant economic


theory for person !.27 All outcomes are explained in a functional
sense by the arguments of gs in equation (9). If we model the ex
post realizations of outcomes, it is entirely reasonable to invoke an
all causes model because ex post all uncertainty has been resolved.
Equation (9) is a production function relating inputs (factors) to
outputs (outcomes). The notation x and us anticipates the econometric problem that some arguments of functional relationship (9)
are observed while other arguments may be unobserved by the analyst. In the analysis of this section, their roles are symmetric.
My notation allows for different unobservables from a common list u to appear in different outcomes.28 gs maps (x, us) into y.
The domain of definition D of gs may differ from the empirical
support. Thus we can think of (9) as mapping logically possible inputs
into logically possible ex post outcomes, but in a real sample we may
observe only a subset of the domain of definition.
A deep structural version of (9) models the variation across
the gs in terms of s as a function of generating characteristics cs that
capture what s is:29
ys gcs ; x; us :

10

The components cs provide the basis for generating the counterfactuals across treatments from a base set of characteristics. This
approach models different treatments as consisting of different bundles of characteristics. g maps c, s, us into y(s), where the domain of
definition D of g may differ from its empirical support. Different
treatments s are characterized by different bundles of the same characteristics that generate all outcomes. This framework provides the
26

This term is discussed in Dawid (2000).


Denote D as the domain of gs : D ! Ry where Ry is the range of y.
28
An alternative notation would use a common u and let gs select out
s-specific components.
29
Now the domain of g, D, is defined for cs, x, us and g : D ! Ry .
27

28

HECKMAN

basis for solving policy problem P3 since new policies (treatments) are
generated as different packages of common characteristics, and all
policies are put on a common basis. If a new policy is characterized by
known transformations of (c, x, us) that lie in the known empirical
support of g, policy forecasting problem P3 can be solved.30 This
point is discussed further in the Appendix.
Part of the a priori specification of a causal model is the choice of
the arguments of the functions gs and g. Analysts may disagree about
appropriate arguments to include based on alternative theoretical frameworks. One benefit of the statistical approach that focuses on problem P1 is that it works solely with the outcomes rather than the inputs.
However, it is silent on how to solve problems P2 and P3 and provides
no basis for interpreting the population-level treatment effects.
Consider alternative models of schooling outcomes of pupils
where s indexes the schooling type (e.g., regular public, charter public,
private secular, and private parochial). The cs are the observed characteristics of schools of type s. The x are the observed characteristics of
the pupil. The us are the unobserved characteristics of both the schools
and the pupil. If we can characterize a proposed new type of school as a
new package of different levels of the same ingredients x, cs, and us and
we can identify (10) over the domain defined by the new package, we
can solve problem P3. If the same schooling input (same cs) is applied
to different students (those with different x) and we can identify (9) or
(10) over the new domain of definition, we solve problem P2. By
digging deeper into the causes of the effects we can do more than
just compare the effects of treatments in place with each other. In
addition, as we shall see, modeling the us and its relationship with the
corresponding unobservables in the treatment choice equation is informative on appropriate identification strategies.
Equations (9) and (10) describing ex post outcomes are sometimes called Marshallian causal functions (see Heckman 2000).
Assuming that the components of (x, us) or (cs, x, us) can be independently varied or are variation-free,31 a feature that may or may not be
30

See Heckman and Vytlacil (2005, 2006a).


The requirement is that if (X ,U) or (C,X ,U) are the domains of
or
(C,X ,U)
(9)
and
(10),
(X ,U) (X 1     X N U 1     U M )
(C1     CK X 1     X N U 1     U M ), where we assume K components in
C, N components in X , and M components in U. This means that we can vary one
variable without necessarily varying another.
31

THE SCIENTIFIC MODEL OF CAUSALITY

29

produced by the relevant theory, we may vary each argument of these


functions to obtain a causal effect of that argument on the outcome.
These thought experiments are for hypotheticals.
Changing one coordinate while fixing the others produces a
Marshallian ceteris paribus causal effect of a change in that coordinate on the variable. Varying cs sets different treatment levels.
Variations in x,us among persons explains why people facing the
same characteristics cs respond differently to the same treatment s.
Variations in us not observed by the analyst explain why people with
the same x values respond differently.
The ceteris paribus variation used to define causal effects need
not be for a single variable of the function. A treatment generally
consists of a package of characteristics and if we vary the package
from cs to cs0 , we get different treatment effects.
I use lowercase notation produced from the theory to denote
fixed values. I use uppercase notation to denote random variables. In
defining equations (9) and (10), I have explicitly worked with fixed
variables that are manipulated in a hypothetical way as in algebra or
elementary physics. In a purely deterministic world, agents would act
on these nonstochastic variables. Even if the world is uncertain,
ex post, after the realization of uncertainty, the outcomes of uncertain
inputs are deterministic. Some components of us may be random
shocks realized after decisions about treatment are made.
Thus if uncertainty is a feature of the environment, equations
(9) and (10) can be interpreted as ex post realizations of the counterfactual as uncertainty is resolved. Ex ante versions of these relationships may be different. From the point of view of agent ! with
information set I ! , the ex ante expected value of Y(s, !) is,32
EYs; ! j I ! EgCs !; X!; Us; ! j I ! ;

11

where Cs, X, Us are random variables generated from a distribution


that depends on the agents information set, indexed by I ! . This
distribution may differ from the distribution produced by reality
32

The expectation might be computed using the information sets of the


relevant decision maker (e.g., the parents in the case of the outcomes of the child)
who might not be the agent whose outcomes are measured. These random
variables are drawn from agent !s subjective distribution.

30

HECKMAN

or nature if agent expectations are different from objective reality.33


In the presence of intrinsic uncertainty, the relevant decision maker
acts on equation (11), but the ex post counterfactual is
Ys; ! EYs; ! j I ! s; !;

12

where (s, !) satisfies E((s, !)jI ! ) 0. In this interpretation, the


information set of agent ! before realizations occur, I ! , is part of
the model specification. This discussion clarifies the distinction
between deterministic (ex post) outcomes and intrinsically random
(ex ante) outcomes discussed in Section 1.
This statement of the basic deterministic model reconciles the all
causes model (9) and (10) with a model of intrinsic uncertainty favored
by some statisticians (see Dawid 2000 and the following discussion).
Ex ante, there is uncertainty at the agent (!) level but ex post there is not.
Realization (s, !) is an ingredient of the ex post all causes model but
not the subjective ex ante all causes model. The probability law used by
the agent to compute the expectation of Cs(!), X(!), Us(!) may differ
from the objective distribution, i.e., the distribution that generates the
observed data. In the ex ante all causes model, manipulations of I !
define the ex ante Marshallian causal parameters.
Thus from the point of view of the agent we can vary elements
in I ! to produce Marshallian ex ante causal response functions. The
ex ante treatment effect from the point of view of the agent for
treatment s and s0 is
EYs; !jI !  EYs0 ; ! j I ! :

13

However, agents may not act on these ex ante effects if they have
decision criteria (utility functions) that are not linear in Y(s, !),
 I discuss ex ante valuations of outcomes in the next
s 1, . . . , S.
section.
The value of the scientific (or explicitly structural) approach to
the construction of counterfactuals is that it explicitly models the
unobservables and the sources of variability among observationally
33
Thus agents do not necessarily use rational expectations, so the distribution used by the agent to make decisions need not equal the distribution
generating the data.

THE SCIENTIFIC MODEL OF CAUSALITY

31

identical people. Since it is the unobservables that give rise to selection


bias and problems of inference that are central to empirically rigorous
causal analysis, analysts using the scientific approach can draw on
scientific theory and in particular choice theory to design and justify
methods to control for selection bias. This avenue is not available to
adherents of the statistical approach. Statistical approaches that are
not explicit about the sources of the unobservables make strong
implicit assumptions which, when carefully exposited, are often unattractive. We exposit some of these assumptions in Section 5.
The models for counterfactualsequations (9)(13)are derived
from theory. The arguments of these functions are varied by hypothetical
manipulations to produce outcomes. These are thought experiments.
When analysts attempt to construct counterfactuals empirically, they
must carefully distinguish between these theoretical relationships and
the empirical relationships determined by the available evidence.
The data used to determine these functions may be limited in
their support. (The support is the region of the domain of definition
where we have data on the function.)34 In this case we cannot fully
identify the theoretical relationships. In addition, in the support, the
components of X, Us and I ! may not be variation-free even if they are
in the hypothetical domain of definition of the function. A good
example is the problem of multicollinearity. If the X in a sample are
linearly dependent, it is not possible to identify the Marshallian causal
function with respect to variations in x over the available support
even if we can imagine hypothetically varying the components of x
over the domains of definition of the functions (9) or (10).
Thus in the available data (i.e., over the empirical support), one of
the X (gender) may be perfectly predictable by the other X. With limited
empirical supports that do not match the domain of definition of the
outcome equations, one may not be able to identify the Marshallian causal
effect of gender even though one can define it in some hypothetical model.
In empirical samples, gender may be predictable in a statistical sense by
other empirical factors. Hollands 1986 claim that the causal effects of
race or gender are meaningless conflates an empirical problem (task 2 in
Table 1) with a problem of theory (task 1 in Table 1). The scientific
34
Thus if Dx is the domain of x, the support of x is the region
Supp (x) Dx such that the data density f (x) satisfies the condition f (x) > 0
for x 2 Supp (x).

32

HECKMAN

approach sharply distinguishes these two issues. One can in theory define
the effect even if one cannot identify it from population or sample data.
I next turn to an important distinction between fixing and
conditioning on factors that gets to the heart of the distinction
between causal models and correlational relationships. This point
is independent of any problem with the supports of the samples
compared to the domains of definition of the functions.

2.2. Fixing Versus Conditioning


The distinction between fixing and conditioning on inputs is central to
distinguishing true causal effects from spurious causal effects. In an
important paper, Haavelmo (1943) made this distinction in linear
equations models. It is the basis for Pearls (2000) book on causality
that generalizes Haavelmos analysis to nonlinear settings. Pearl defines
an operator do to represent the mental act of fixing a variable to
distinguish it from the action of conditioning which is a statistical
operation. If the conditioning set is sufficiently rich, fixing and conditioning are the same in an ex post all causes model.35 Pearl suggests a
particular physical mechanism for fixing variables and operationalizing
causality, but it is not central to his or any other definition of causality.
Pearls analysis conflates the three tasks of Table 1.
An example of fixing versus conditioning is most easily illustrated in a linear regression model of the type analyzed by Haavelmo
(1943). Let y x u. Although both y and u are scalars, x may be
a vector. The linear equation maps (x, u) into y: (x, u) 7! y. Suppose
that the support of random variable (X, U) in the data is the same as
the domain of (x, u) that are fixed in the hypothetical thought experiment and that the (x, u) are variation-free (i.e., they can be independently varied coordinate by coordinate). Thus we abstract from the
problem of limited support that is discussed in the preceding section.
We may write (dropping the ! notation for random variables)
Y X U:

35

Florens and Heckman (2003) carefully distinguish conditioning from


fixing, and generalize Pearls analysis to both static and dynamic settings.

THE SCIENTIFIC MODEL OF CAUSALITY

33

Here nature or the real world picks (X, U) to determine Y.


X is observed by the analyst and U is not observed, and (X, U) are
random variables. This is an all causes model in which (X, U) 7! Y.
The variation generated by the hypothetical model varies one coordinate of (X, U), fixing all other coordinates to produce the effect of the
variation on the outcome Y. Nature (as opposed to the model) may
not permit such variation.
Formally, we can write this model formulated at the population level as a conditional expectation,
EYjX; U X U:
Since we condition on both X and U, there is no further source
of variation in Y. This is a deterministic model that coincides with
the all causes model. Thus on the support, which is also assumed to
be the domain of definition of the function, this model is the
same model as the deterministic, hypothetical model, y x u.
Fixing X at different values corresponds to doing different thought
experiments with the X. Fixing and conditioning are the same in this
case.
If, however, we only condition on X in the sample, we obtain
EYjX X EUjX:36

14

This relationship does not generate U-constant (Y, X) relationships. It


generates only an X-constant relationship. Unless we condition on all
of the causes (the right hand side variables), the empirical relationship (14) does not identify causal effects of X on Y. The variation in
X also moves the conditional mean of U unless U is independent of X.
This analysis readily generalizes to a general nonlinear model
y g (c, x, u). A model specified in terms of random variables C, X,
U with the same support as c, x, u has as its conditional expectation
g(C, X, U) under general conditions. Conditioning only on C, X does
not in principle identify g(c, x, u) or any of its derivatives (if they
exist) or differences of outcomes defined in terms of c and x.

36

I assume that the mean of U is finite.

34

HECKMAN

Conditioning and fixing on the arguments of g or gs are the


same in an all causes model if all causes are accounted for.
Otherwise, they are not the same. This analysis can be generalized
to account for the temporal resolution of uncertainty if we include
 (s, !) as an argument in the ex post causal model. The outcomes
can include both objective outcomes Y(s, !) and subjective outcomes
V(Y(s, !), !).
Statisticians and epidemiologists have great difficulty with the
distinction between fixing and conditioning because they typically
define the models they analyze in terms of some type of conditioning.
However, thought experiments in models of hypotheticals that vary
factors are distinct from variations in conditioning variables that
conflate the effects of variation in X, holding U fixed, with the effects
of X in predicting the unobserved factors (the U) in the outcome
equations.
2.3. Modeling the Choice of Treatment
Parallel to the models for outcomes are models for the choice of
treatment. Consider ex ante personal valuations of outcomes based
on expectations of gains from receiving treatment s:
EVYs; !; Ps; !; Cs !; !jI ! ; s 2 S;
where P(s, !) is the price or cost the agent must pay for participation
in treatment s. We write P(s, !) K(Z(s, !), (s, !)). I allow utility V
to be defined over the characteristics that generate the treatment
outcome (e.g., quality of teachers in a schooling choice model) as
well as other attributes of the consumer. In parallel with the gs
function generating the Y(s, !), we write
VYs; !; Ps; !; Cs !; ! fYs; !; Zs; !; Cs !; s; !; !:
Parallel to the analysis of outcomes, we may keep Cs(!) implicit and
use fs functions instead of f.
My analysis includes both measured and unmeasured
attributes. The agent computes expectations against his/her subjective
distribution of information. I allow for imperfect information
by postulating an !-specific information set. If agents know all

THE SCIENTIFIC MODEL OF CAUSALITY

35

components of future outcomes, the uppercase letters become lowercase variables that are known constants. The I ! are the causal factors
for !. In a utility-maximizing framework, choice b
s is made if b
s is
maximal in the set of valuations of potential outcomes:
fEVYs; !; Ps; !; Cs !; !jI !  : s 2 Sg:
In this interpretation, the information set plays a key role in specifying
agent preferences. Actual realizations may not be known at the time
decisions are made. Accounting for uncertainty and subjective valuations of outcomes (e.g., pain and suffering for a medical treatment) is a
major contribution of the scientific approach. The factors that lead an
agent to participate in treatment s may be dependent on the factors
affecting outcomes. Modeling this dependence is a major source of
information used in the scientific approach to constructing counterfactuals from real data, as I demonstrate in Section 4. A parallel
analysis can be made if the decision maker is not the same as the
agent whose objective outcomes are being evaluated.
2.4. The Scientific Model Versus the NeymanRubin Model
Many statisticians and social scientists invoke a model of counterfactuals and causality attributed to Donald Rubin by Paul Holland (1986)
but which actually dates back to Neyman (1923).37 Neyman and Rubin
postulate counterfactuals fY(s, !)gs2S without modeling the factors
determining the Y(s, !) as I have done in equations (9)(12), using
the scientific, structural approach. Rubin and Neyman offer no model
of the choice of which outcome is selected. Thus there no lowercase,
all causes models explicitly specified in this approach, nor is there any
discussion of the science or theory producing the outcomes studied.
In my notation, Rubin assumes (A-1) and (A-2) as presented in
Section 1.38 Recall that (A-1) assumes no general equilibrium effects or
social interactions among agents. Thus the outcome for the person is the
37

The framework attributed to Rubin was developed in statistics by


Neyman (1923), Cox (1958), and others. Parallel frameworks were independently
developed in psychometrics (Thurstone 1930) and economics (Haavelmo 1943;
Roy 1951; Quandt 1958, 1972).
38
Rubin (1986) calls these two assumptions SUTVA for Stable Unit
Treatment Value Assumption.

36

HECKMAN

same whether one person receives treatment or many receive treatment.


(A-2) says that however ! receives s, the same outcome arises. (A-2) also
rules out randomization bias where the act of randomization affects the
potential outcomes.39
More formally, the Rubin model assumes the following:
R-1 fY(s, !)gs2S , a set of counterfactuals defined
for ex post outcomes (no valuations of outcomes or
specification of treatment selection rules).
R-2 (A-1) (No social interactions).
R-3 (A-2) (Invariance of counterfactual to assignment
mechanism of treatment).
R-4 P1 is the only problem of interest.
R-5 Mean causal effects are the only objects of interest.
R-6 There is no simultaneity in causal effects, i.e.,
outcomes cannot cause each other reciprocally (see
Holland 1988).
The scientific model (1) decomposes the Y(s, !), s 2 S into its
determinants; (2) considers valuation of outcomes as an essential
ingredient of any study of causal inference; (3) models the choice of
treatment and uses choice data to infer subjective valuations of treatment; (4) uses the relationship between outcomes and treatment
choice equations to motivate, justify, and interpret alternative identifying strategies; (5) explicitly accounts for the arrival of information
through ex ante and ex post analyses; (6) considers distributional
causal parameters as well as mean effects; (7) addresses problems
P1P3; (8) allows for nonrecursive (simultaneous) causal models.
I develop nonrecursive models in the next section.
In the NeymanRubin model, the sources of variability generating Y(s, !) as a random variable are not specified. The causal
effect of s compared to s0 is defined as the treatment effect in
equation (1). Holland (1986, 1988) argues that it is an advantage of
the Rubin model that it is not explicit about the sources of variability
among observationally identical people, or about the factors that

39

See Heckman (1992) or Heckman, LaLonde, and Smith (1999) for


discussions and evidence on this question.

THE SCIENTIFIC MODEL OF CAUSALITY

37

generate Y(s, !). Holland and Rubin focus on mean treatment effects
as the interesting causal parameters.
The scientific (econometric) approach to causal inference supplements the model of counterfactuals with models of the choice of
counterfactuals fD(s, !)gs2S generated by the maps p (!) and the
relationship between choice equations and the counterfactuals. The
D(s, !) are assumed to be generated by the collection of random
variables (Cs (!), Z(s, !), (s, !), Y(s, !)jI ! ), s 2 S, where Cs(!) is
the characteristic of the treatment s for person !, Z(s, !) are observed
determinants of costs, the (s, !) are unobserved (by the analyst) cost
(or preference) factors and Y(s, !) are the outcomes, and the j
denotes that these variables are defined conditional on I ! (the agents
information set).40 Along with the ex ante valuations that generate
D(s, !) are the ex post valuations discussed in Section 1.6.
Random utility models generating D(s, !) go back to
Thurstone (1930) and McFadden (1974, 1981).41 The full set of counterfactual outcomes for each agent is assumed to be unobserved by
the analyst. It is the dependence of unmeasured determinants of
treatment choices with unmeasured determinants of potential outcomes that gives rise to selection bias in empirically constructing
counterfactuals and treatment effects, even after conditioning on the
observables. Knowledge of the relationship between choices and
counterfactuals suggests appropriate methods for solving selection
problems. By analyzing the relationship of the unobservables in the
outcome equation, and the unobservables in the treatment choice
equation, the analyst can use a priori theory to devise appropriate
estimators to identify causal effects.
The scientific approach is more general than the Neyman
Rubin model because it emphasizes the welfare of the agents being
studied (through VG or V(Y(s, !), !))the subjective evaluations
as well as the objective evaluations. The econometric approach also

40
If other agents make the treatment assignment decisions, then the
determinants of D(s; !) are modified according to what is in their information set.
41
Corresponding to these random variables are the deterministic all causes
counterparts d(s), cs, z(s), (s), {y(s)}, i, where the (fz(s)gs2S ; fcs gs2S ; f(s)gs2S ;
fy(s)gs2S ; i) generate the d(s) 1 if (fz(s)gs2S ; fcs gs2S ; f(s)gs2S ; fy(s)gs2S ) 2 , a
subset of the domain of the generators of d (s). Again the domain of definition of
d(s) is not necessarily the support of z(s; !); cs (!); (s; !); fY(s; !)gs2S and I ! .

38

HECKMAN

distinguishes ex ante from ex post subjective evaluations, so it can


measure both agent satisfaction and regret.42
In addition, modelling Y(s, !) in terms of characteristics of treatment, and of the treated, facilitates comparisons of counterfactuals and
derived causal effects across studies where the composition of programs
and treatment group members may vary. It also facilitates the construction of counterfactuals on new populations and the construction of
counterfactuals for new policies. The NeymanRubin framework focuses
exclusively on population-level mean causal effects or treatment effects
for policies actually experienced and provides no framework for extrapolation of findings to new environments or for forecasting new policies
(problems P2 and P3). Its focus on population mean treatment effects
elevates randomization and matching to the status of preferred estimators. Such methods cannot identify distributions of treatment effects or
general quantiles of treatment effects.
Another feature of the NeymanRubin model is that it is
recursive. It cannot model causal effects of outcomes that occur
simultaneously. I now present a model of simultaneous causality.
2.5. Nonrecursive (Simultaneous) Models of Causality
A system of linear simultaneous equations captures interdependence
among outcomes Y. For simplicity, I focus on ex post outcomes so I ignore
the revelation of information over time. To focus on the main ideas of this
section, I assume that the domain of definition of the model is the same as
the support of the population data. Thus the model for values of uppercase
variables has the same support as the domain of definition for the model in
terms of lowercase variables.43 The model developed in this section is rich
enough to model interactions among agents.44 I write this model in terms
of parameters ( , B), observables (Y, X), and unobservables U as
Y BX U;

EU 0;

15

42
See Cunha, Heckman, and Navarro (2005a,b) for estimates of subjective evaluations and regret in schooling choices.
43
This approach merges tasks 1 and 2 in Table 1. I do this here because
the familiarity of the simultaneous equations model as a statistical model makes
the all causes ex post version confusing to many readers familiar with this model.
44
For simplicity, I work with the linear model in the text, developing the
nonlinear case in footnotes.

THE SCIENTIFIC MODEL OF CAUSALITY

39

where Y is now a vector of endogenous and interdependent variables, X is


exogenous (E(UjX) 0), and is a full rank matrix. A better nomenclature, suggested by Leamer (1985), is that the Y are internal variables
determined by the model and the X are external variables specified outside
the model.45 This definition distinguishes two issues: (1) defining variables
(Y) that are determined from inputs outside the model (the X) and (2)
determining the relationship between observables and unobservables.46
When the model is of full rank ( 1 exists), it is said to be complete. A
complete model produces a unique Y from a given (X, U). A complete model
is said to be in reduced form when equation (15) is multiplied by 1.
The reduced form is Y X R where   1B and R 1U.47
This is a linear-in-parameters all causes model for vector Y, where the
causes are X and R. The structure is ( , B), U, where U is the
variance-covariance matrix of U. The reduced form slope coefficients are
, and R is the variance-covariance matrix of R.48 In the population
generating (15), least squares recovers  provided X, the variance of X,
is nonsingular (no multicollinearity). In this linear-in-parameters equation
setting, the full rank condition for X is a variation-free condition on the
external variables. The reduced form solves out for the dependence among
the Y. The linear-in-parameters model is traditional. Nonlinear versions
are available (Fisher 1966; Matzkin 2004).49 For simplicity, I stick to the
linear version, developing the nonlinear version in footnotes.
The structural form (15) is an all causes model that relates in a
deterministic way outcomes (internal variables) to other outcomes
(internal variables) and external variables (the X and U). Without
some restrictions, certain ceteris paribus manipulations associated
45
This formulation is static. In a dynamic framework, Yt would be the
internal variables and the lagged Y, Yt  k, k > 0, would be external to period t
and be included in the Xt. Thus we could work with lagged dependent variables.
The system would be Yt BXt Ut, E(Ut) 0.
46
In a time-series model, the internal variables are Yt determined in
period t.
47
In this section only,  refers to the reduced form coefficient matrix
and not the set of policies p, as in earlier sections.
48
The original formulations of this model assumed normality so
that only means and variances were needed to describe the joint distributions of
(Y, X).
49
The underlying all causes model writes y Bx u, y x r and
1
  B, r 1u. Recall that I assume that the domain of the all causes
model is the same as the support of (x, u). Thus there is a close correspondence
between these two models.

40

HECKMAN

with the effect of some components of Y on other components of Y


are not possible within the model. I now demonstrate this point.
For specificity, consider a two-person model of social interactions.
Y1 is the outcome for person 1; Y2 is the outcome for person 2. This could
be a model of interdependent consumption where the consumption of
person 1 depends on the consumption of person 2 and other person-1specific variables (and possibly other person-2-specific variables). It could
also be a model of test scores. We can imagine populations of data
generated from sampling the same two-person interaction over time or
sampling different two-person couplings at a point in time.
Assuming that the preferences are interdependent, we may write
Y1 a1 12 Y2 11 X1 12 X2 U1

16a

Y2 a2 21 Y1 21 X1 22 X2 U2 :

16b

This model is sufficiently flexible to capture the notion that the consumption of person 1 (Y1) depends on the consumption of person 2 (if
6 0), as well as person 1s value of X (if  11
6 0), X1 (assumed to be
12
observed), person 2s value of X, X2 (if 12 0), and unobservable
factors that affect person 1 (U1). The determinants of person 2s consumption are defined symmetrically. I allow U1 and U2 to be freely
correlated. I assume that U1 and U2 are mean independent of (X1, X2) so
EU1 jX1 ; X2 0

17a

EU2 jX1 ; X2 0:

17b

and

Completeness guarantees that (16a) and (16b) have a determinate solution for (Y1,Y2).
Applying Haavelmos argument to (16a) and (16b), the causal
effect of Y2 on Y1 is 12. This is the effect on Y1 of fixing Y2 at different
values, holding constant the other variables in the equation.
Symmetrically, the causal effect of Y1 on Y2 is g 21. Conditioning,that
is, using least squareswhich is the method of matching, in general fails
to identify these causal effects because U1 and U2 are correlated with Y1
and Y2. This is a traditional argument. It is based on the correlation
between Y2 and U1. But even if U1 0 and U2 0, so that there are no

41

THE SCIENTIFIC MODEL OF CAUSALITY

unobservables, matching or least squares breaks down because Y2 is


perfectly predictable by X1 and X2. We cannot simultaneously vary Y2,
X1, and X2. This is the essence of the problem of defining a causal effect.
To see why, we derive the reduced form of this model.
Assuming completeness, the reduced form outcomes of the
model after social interactions are solved out can be written as
Y1 10 11 X1 12 X2 R1

18a

Y2 20 21 X1 22 X2 R2 :

18b

Least squares (matching) can identify the ceteris paribus effects of X1


and X2 on Y1 and Y2 because E(R1jX1, X2) 0 and E(R2jX1, X2) 0.
Simple algebra informs us that
11 21 21
1  12 21
21 11 21

1  12 21

12 22 12
1  12 21
:
12 12 22

1  12 21

11

12

21

22

19

U1 21 U2
1  12 21
:
12 U1 U2
R2
1  12 21

and

R1

Observe that because R2 depends on both U1 and U2 in the general case,


Y2 is correlated with U1 (through the direct channel of U1 and through
the correlation between U1 and U2). Without any further information on
the variances of (U1, U2) and their relationship to the causal parameters,
we cannot isolate the causal effects g 12 and g 21 from the reduced form
regression coefficients. This is so because holding X1, X2, U1, and U2
fixed in (16a) or (16b), it is not in principle possible to vary Y2 or Y1,
respectively, because they are exact functions of X1, X2, U1, and U2.
This exact dependence holds true even if U1 0 and U2 0 so
that there are no unobservables.50 In this case, which is thought to be the
most favorable to the application of least squares or matching to (16a)
and (16b), it is evident from (18a) and (18b) that when R1 0 and

50

See Fisher (1966).

42

HECKMAN

R2 0, Y1 and Y2 are exact functions of X1 and X2. There is no mechanism yet specified within the model to independently vary the right-hand
sides of equations (16a) and (16b).51 The X effects on Y1 and Y2, identified through the reduced forms, combine the direct effects (through  ij)
and the indirect effects (as they operate through Y1 and Y2, respectively).
If we assume exclusions ( 12 0) or (21 0) or both, we can
identify the ceteris paribus causal effects of Y2 on Y1 and of Y1 on Y2
respectively. Thus if 12 0 from the reduced form,
12
12 :
22
If  21 0, we obtain
21
21 :
11
These exclusions say that the social interactions only operate through
the Ys. Person 1s consumption depends only on person 2s consumption and not on his or her X2 or directly through his or her U2. Person
2 is modeled symmetrically versus person 1. Observe that I have not
ruled out correlation between U1 and U2. When the procedure for
identifying causal effects is applied to samples, it is called indirect
least squares. The method traces back to Haavelmo (1943, 1944).52
The intuition for these results is that if  12 0, we can vary Y2
in equation (16a) by varying the X2. Since X2 does not appear in the
51
Some readers of an earlier draft of this paper suggested that the mere
fact that we can write (16a) and (16b) means that we can imagine independent
variation. By the same token, we can imagine a model

Y 0 1 X1 2 X2 ;
but if part of the model is (*) X1 X2, the rules of the model constrain X1 X2.
No causal effect of X1 holding X2 constant is possible. If we break restriction (*)
and permit independent variation in X1 and X2, we can define the causal effect
of X1 holding X2 constant.
52
The analysis for social interactions in this section is of independent
interest. It can be generalized to the analysis of N person interactions if the
outcomes are continuous variables. For binary outcomes variables, the same
analysis goes through for the special case analyzed by Heckman and MaCurdy
(1985). However, in the general case, for discrete outcomes generated by latent
variables it is necessary to modify the system to obtain a coherent probability
model; see Heckman (1978).

THE SCIENTIFIC MODEL OF CAUSALITY

43

equation, under exclusion, we can keep U1, X, fixed and vary Y2 using
X2 in (18b) if 22 6 0.53 Symmetrically, by excluding X1 from (16b),
we can vary Y1, holding X2 and U2 constant. These results are more
clearly seen when U1 0 and U2 0.
Observe that in the model under consideration, where the
domain of definition and the supports of the variables coincide, the
causal effects of simultaneous interactions are defined if the parameters are identified in the traditional Cowles definition of identification (e.g., see Ruud 2000 for a modern discussion of these conditions).
A hypothetical thought experiment justifies these exclusions. If agents
do not know or act on the other agents X, these exclusions are
plausible.
An implicit assumption in using (16a) and (16b) for causal
analysis is invariance of the parameters ( , , U) to manipulations
of the external variables. This invariance embodies the key idea in
assumption (A-2). Invariance of the coefficients of equations to classes
of manipulation of the variables is an essential part of the definition of
structural models that I develop more formally in the next section.
This definition of causal effects in an interdependent system
generalizes the recursive definitions of causality featured in the statistical treatment effect literature (Holland 1988; Pearl 2000). The key to
this definition is manipulation of external inputs and exclusion, not
randomization or matching. Indeed matching or, equivalently, OLS,
using the right-hand side variables of (16a) and (16b), does not
identify causal effects as Haavelmo (1943) established long ago. We
can use the population simultaneous equations model to define the
class of admissible variations and address problems of definitions
(task 1 in Table 1). If for a given model, the parameters of (16a) or
(16b) shift when external variables are manipulated, or if external
variables cannot be independently manipulated, causal effects of one
internal variable on another cannot be defined within that model. If
people were randomly assigned to pair with their neighbors, and the
parameters of (16a) were not affected by the randomization, then Y2
would be exogenous in equation (16b) and we could identify causal
53

Notice that we could also use U2 as a source of variation in (18b) to


shift Y2. The roles of U2 and X2 are symmetric. However, if U1 and U2 are
correlated, shifting U2 shifts U1 unless we control for it. The component of U2
uncorrelated with U1 plays the role of X2.

44

HECKMAN

effects by least squares. At issue is whether such a randomization


would recover 12. It might fundamentally alter agent 1s response to
Y2 if that person is randomly assigned as opposed to being selected by
the agent. Judging the suitability of an invariance assumption entails
a thought experimenta purely mental act.
Controlled variation in external forcing variables is the key to
defining causal effects in nonrecursive models. It is of some interest to
readers of Pearl (2000) to compare my use of the standard simultaneous equations model of econometrics in defining causal parameters
to his. In the context of equations (16a) and (16b), Pearl defines a
causal effect by shutting one equation down or performing surgery in his colorful language.
He implicitly assumes that surgery, or shutting down an
equation in a system of simultaneous equations, uniquely fixes one
outcome or internal variable (the consumption of the other person in
my example). In general, it does not. Putting a constraint on one
equation places a restriction on the entire set of internal variables. In
general, no single equation in a system of simultaneous equations
uniquely determines any single outcome variable. Shutting down
one equation might also affect the parameters of the other equations
in the system and violate the requirements of parameter stability.
A clearer manipulation is to assume that it is possible to fix Y2
by setting 12 0. Assume that U1 and U2 are uncorrelated.54 This
makes the model recursive. It assumes that person 1 is unaffected by
the consumption of person 2. Under these assumptions, we can
regress Y1 on Y2, X1, and X2 in the population and recover all of
the causal parameters of (16a). Variation in U2 breaks the perfect
collinearity among Y2, X1, and X2. It is far from obvious, however,
that one can freely set parameters without affecting the rest of the
parameters of the model.
Shutting down an equation or fiddling with the parameters in
is not required to define causality in an interdependent, nonrecursive
system or to identify causal parameters. The more basic idea is exclusion of different external variables from different equations which,
when manipulated, allow the analyst to construct the desired causal
quantities.
54

Alternatively, we can assume that it is possible to measure U1 and


control for it.

THE SCIENTIFIC MODEL OF CAUSALITY

45

One can move from the problem of definition (task 1 in Table 1)


to identification (task 2) by using population analog estimation methodsin this case the method of indirect least squares.55 There are
many ways other than through exclusions of variables to identify this
and more general systems. Fisher (1966) presents a general analysis of
identification in both linear and nonlinear simultaneous equations
systems. Matzkin (2004) is a recent substantial extension of this
literature.
In the context of the basic nonrecursive model, there are many
possible causal variations, richer than what can be obtained from the
reduced form. Using the reduced form (Y X R), we can define
causal effects as ceteris paribus effects of variables in X or R on Y.
This definition solves out for all of the intermediate effects of the
internal variables on each other. Using the structure in equation (15),
we can define the effect of one internal variable on another holding
constant the remaining internal variables and (X, U). It has just been
established that such causal effects may not be defined within the
rules specified for a particular structural model. Exclusions and other
restrictions discussed in Fisher (1966) make definitions of causal
effects possible under certain conditions.
One can, in general, solve out from the general system of
equations for subsets of the Y (e.g., Y* where Y (Y*, Y**)) using
the reduced form of the model and use quasi-structural models
to define a variety of causal effects that solve out for some but not
all of the possible causal effects of Y on each other. These quasistructural models may be written as
** Y** ** X U** :
This expression is obtained by using the reduced form for component
Y*:Y* *X R* and substituting for Y* in (15). U** is the error
term associated with this representation. There are many possible quasistructural models. Causal effects of internal variables may or may not be
defined within them, depending on the assumed a priori information.
The causal effect of one component of Y** on another does not
fix Y* but allows the Y* components to adjust as the components of
Y** and the X are varied. Thus the Y* are not being held fixed when
55

Two-stage least squares would work as well.

46

HECKMAN

X and/or components of the Y** are varied. Viewed in this way, the
reduced form and the whole class of quasi-structural models do not
define any ceteris paribus causal effect relative to all of the variables
(internal and external) in the system since they do not fix the levels of
the other Y or Y* in the case of the quasi-structural models.
Nonetheless, the reduced form may provide a good guide to forecasting the effects of certain interventions that affect the external variables. The quasi-structural models may also provide a useful guide for
predicting certain interventions, where Y** are fixed by policy. The
reduced form defines a net causal effect of variations in X as they
affect the internal variables. There are many quasi-structural models
and corresponding thought experiments.
This discussion demonstrates another reason why causal knowledge is provisional. Different analysts may choose different subsystems
of equations derived from equation (15) to work with and define
different causal effects within the different possible subsystems. Some
of these causal effects may not be identified, while others may be.
Systems smaller or larger than (15) can be imagined. The role of a
priori theory is to limit the class of models and the resulting class of
counterfactuals and to define which ones are interesting.
I now present a basic definition of structure in terms of invariance of equations to classes of interventions. Invariance is a central
idea in causal analysis and in policy analysis.
2.6. Structure as Invariance
A basic definition of a system of structural relationships is that it is a
system of equations invariant to a class of modifications or interventions. In the context of policy analysis, this means a class of policy
modifications. This is the definition that was proposed by Hurwicz
(1962). It is implicit in Marschak (1953) and it is explicitly utilized by
Sims (1977), Lucas and Sargent (1981), and Leamer (1985), among
others. This definition requires a precise definition of a policy, a class
of policy modifications, and specification of a mechanism through
which policy operates.
The mechanisms generating counterfactuals and the choices of
counterfactuals have already been characterized in Sections 2.1 and
2.3. Policies can act on preferences and the arguments of preferences
(and hence choices), on outcomes Y(s, !) and the determinants

THE SCIENTIFIC MODEL OF CAUSALITY

47

affecting outcomes or on the information facing agents. Recall that gs,


s 2 S, generates outcomes while fs, s 2 S, generates evaluations.
Specifically,
1.

2.

3.

Policies can shift the distributions of the determinants of


outcomes and choices (C, Z, X, U, ), where C fCs (!)gs2S ,
Z fZ(s, !)gs2S ,  f(s, !)gs2S and U fUs (!)gs2S in the
population. This may entail defining the gs and fs over new
domains. Let Q (C, Z, X, U, ). Policies shifting the distributions of these variables are characterized by maps TQ : Q7!Q0 .
Policies may select new f, g or ffs , gs gs2S functions.56 In particular, new arguments (e.g., amenities or characteristics of programs)
may be introduced as a result of policy actions creating new
attributes. Policies shifting functions map f, g or ffs , gs gs2S into
new functions Tf : fs 7!f0s ; Tg : gs 7!g0s . This may entail changes in
functional forms with a stable set of arguments as well as changes
in arguments of functions.
Policies may affect individual information sets (I ! )!2 :TI ! :
I ! 7!I 0! .

Clearly, any particular policy may incorporate elements of all three


types of policy shifts.
Parameters of a model or parameters derived from a model are
said to be policy invariant if they are not changed (are invariant)
when policies are implemented. This notion is partially embodied in
assumption (A-2), which is defined solely in terms of ex post outcomes. More generally, policy invariance for f,g or ffs , gs gs2S requires
the following:
(A-3) The functions f,g or {fs, gs}s2S are the same for
all values of the arguments in their domain of definition no matter how their arguments are determined.
This definition can be made separately for f, g, fs, gs or any function
derived from them. It requires that when we change an argument of a
function it does not matter how we change it.

56

By fs, we mean s-specific valuation functions.

48

HECKMAN

In the simultaneous equations model analyzed in the last section, invariance requires stability of , B, and U to interventions.
Such models can be used to accurately forecast the effects of policies
that can be cast as variations in the inputs to the model. Policyinvariant parameters are not necessarily causal parameters, as we
noted in our analysis of reduced forms in the preceding section.
Thus, in the simultaneous equations model, depending on the a priori
information available, no causal effect of one internal variable on
another may be defined but if  is invariant to modifications in X,
the reduced form is policy invariant for those modifications. The class
of policy-invariant parameters is thus distinct from the class of causal
parameters, but invariance is an essential attribute of a causal model.
For counterfactuals Y(s, !), if assumption (A-3) is not postulated, all
of the treatment effects defined in Section 1 would be affected by
policy shifts. Rubins assumption (A-2) makes Y(s, !) invariant to
policies that change f but not policies that change g or the support of
Q. Within the treatment effects framework, a policy that adds a new
treatment to S is not policy invariant for treatment parameters comparing the new treatment to any other treatment unless the analyst
can model all policies in terms of a generating set of common characteristics specified at different levels. The lack of policy invariance
makes it difficult to forecast the effects of new policies using treatment
effect models within the framework of the Appendix.
Deep structural parameters generating the f and g are invariant to policy modifications that affect technology, constraints, and
information sets except when the policies extend the historical supports. Invariance can only be defined relative to a class of modifications and a postulated set of preferences, technology, constraints, and
information sets. Thus causal parameters can be precisely identified
only within a class of modifications.
2.7. Marschaks Maxim and the Relationship Between Structural
Literature and Statistical Treatment Effect Literature
The absence of explicit models is a prominent feature of the statistical
treatment effect literature. Scientifically well-posed models make
explicit the assumptions used by analysts regarding preferences, technology, the information available to agents, the constraints under
which they operate, and the rules of interaction among agents in

THE SCIENTIFIC MODEL OF CAUSALITY

49

market and social settings and the sources of variability among persons. These explicit features make these models, like all scientific
models, useful vehicles: (1) for interpreting empirical evidence using
theory; (2) for collating and synthesizing evidence using theory; (3) for
measuring the welfare effects of policies; and (4) for forecasting the
welfare and direct effects of previously implemented policies in new
environments and the effects of new policies.
These features are absent from the modern treatment effect
literature. At the same time, this literature makes fewer statistical
assumptions in terms of exogeneity, functional form, exclusion, and
distributional assumptions than the standard structural estimation
literature in econometrics. These are the attractive features of this
approach.
In reconciling these two literatures, I reach back to a neglected
but important paper by Jacob Marschak. Marschak (1953) noted that
for many specific questions of policy analysis, it is unnecessary to
identify full structural models, where by structural I mean parameters
invariant to classes of policy modifications as defined in the last
section. All that is required are combinations of subsets of the structural parameters, corresponding to the parameters required to forecast particular policy modifications, which are much easier to
identify (i.e., require fewer and weaker assumptions). Thus in the
simultaneous equations system examples, policies that only affect X
may be forecast using reduced forms, not knowing the full structure,
provided that the reduced forms are invariant to the modifications.57 Forecasting other policies may require only partial knowledge of the system. I call this principle Marschaks maxim in honor
of this insight. I interpret the modern statistical treatment effect
literature as implicitly implementing Marschaks maxim where the
policies analyzed are the treatments and the goal of policy analysis
is restricted to evaluating policies in place (task 1; P1) and not in
forecasting the effects of new policies or the effects of old policies
on new environments.
Population mean treatment parameters are often identified under weaker conditions than are traditionally assumed in
econometric structural analysis. Thus to identify the average
57

Thus we require that the reduced form  does not change when we
change the X.

50

HECKMAN

treatment effect for s and s0 we require only E(Y(s, !) j X x) 


E(Y(s0 , !) j X s). We do not have to know the full functional
form of the generating gs functions nor does X have to be
exogenous. The treatment effects may, or may not, be causal
parameters depending on what else is assumed about the model.
Considerable progress has been made in relaxing the parametric structure assumed in the early structural models in econometrics (see Matzkin 2006). As the treatment effect literature is
extended to address the more general set of policy forecasting problems entertained in the structural literature, the distinction between
the two literatures will vanish although it is currently very sharp.
Heckman and Vytlacil (2005, 2006a,b) and Heckman (2006) are
attempts to bridge this gulf.
Up to this point in the essay, everything that has been discussed precisely is purely conceptual, although I have alluded to
empirical problems and problems of identification going from data
of various forms to conceptual models. Models are conceptual and so
are the treatment effects derived from them. The act of defining a
model is distinct from identifying it or estimating it although statisticians often conflate these distinct issues. I now discuss the identification problem, which must be solved if causal models are to be
empirically operational.

3. IDENTIFICATION PROBLEMS: DETERMINING MODELS


FROM DATA
Unobserved counterfactuals are the source of the problems considered in this paper. For a person in state s, we observe Y(s, !) but not
Y(s0 , !), s0 6 s. A central problem in the literature on causal inference
is how to identify counterfactuals and the derived treatment parameters. Unobservables, including missing data, are at the heart of the
identification problem considered here.
Estimators differ in the amount of knowledge they assume that
the analyst has relative to what the agents being studied have when
making their program enrollment decisions (or their decisions are
made for them as a parent for a child). This is strictly a matter of
the quality of the available data. Unless the analyst has access to all of
the relevant information that produces the dependence between

THE SCIENTIFIC MODEL OF CAUSALITY

51

outcomes and treatment rules (i.e., that produces selection bias), he or


she must devise methods to control for the unobserved components of
relevant information. Heckman and Vytlacil (2006b) and Heckman
and Navarro (2004) define relevant information precisely. Relevant
information is the information which, if available to the analyst and
conditioned on, would eliminate selection bias. Intuitively, there may
be a lot of information known to the agent but not known to the
observing analyst that is irrelevant in creating the dependence
between outcomes and choices. It is the information that gives rise
to the dependence between outcomes and treatment choices that
matters for eliminating selection bias.
A priori one might think that the analyst knows a lot less than
the agent whose behavior is being analyzed. At issue is whether the
analyst knows less relevant information, which is not so obvious, if
only because the analyst can observe the outcomes of decisions in a
way that agents making decisions cannot. This access to ex post
information can sometimes give the analyst a leg up on the information available to the agent.
The policy forecasting problems P2 and P3 raise the additional
issue that the support over which treatment parameters and counterfactuals are identified may not correspond to the support to which the
analyst seeks to apply them. Common to all scientific models, there is
the additional issue of how to select (X, Z), the conditioning variables,
and how to deal with them if they are endogenous. Finally, there is the
problem of lack of knowledge of functional forms of the models.
Different econometric methods solve these problems in different
ways. I now present a precise discussion of identification.
3.1. The Identification Problem
The identification problem asks whether theoretical constructs have
any empirical content in a hypothetical population or in real samples.
This formulation considers tasks 2 and 3 in Table 1 together, although
some analysts like to separate these issues, focusing solely on task 2.
The identification problem considers what particular models within a
broader class of models are consistent with a given set of data or facts.
Specifically, we can consider a model space M. This is the set of
admissible models that are produced by some theory for generating
counterfactuals. Elements m 2 M are admissible theoretical models.

52

HECKMAN

We may be interested in only some features of a model. For


example, we may have a rich model of counterfactuals {Y(s, !)}s 2 S ,
but we may be interested in only the average treatment effect
E![Y(s, !)  Y(s0 , !)]. Let the objects of interest be t 2 T, where t
stands for the targetthe goal of the analysis. The target space
T may be the whole model space M or something derived from it.
Define map g: M T. This maps an element m 2 M into an
element t 2 T. In the example in the preceding paragraph, T is the
space of all average treatment effects produced by the models of
counterfactuals. I assume that g is into.58 Associated with each
model is an element t derived from the model, which could be the
entire model itself. Many models may map into the same t so the
inverse map (g1), mapping T to M, may not be well-defined. Thus
many different models may produce the same average treatment
effect.
Let the class of possible information or data be I. Define a map
h: M I. For an element i 2 I, which is a given set of data, there may
be one or more models m consistent with i. If i can be mapped only
into a single m, the model is exactly identified.59 If there are multiple
ms, consistent with i, these models are not identified. Thus, in Figure 1,
many models (elements of M) may be consistent with the same data
(single element of I).
Let M h (i) be the set of models consistent with i.
M h (i) h 1 ({i}) {m 2 M : h(m) i}. The data i reject the other
models M\Mh(i), but are consistent with all models in Mh(i). If Mh(i)
contains more than one element, the data produce set-valued instead
of point-valued identification. If Mh(i) , the empty set, no

58

By this, we mean that for every t 2 T, there is an element m 2 M such


that g sends m to t, i.e., the image of g is the entire set T. Of course, g may send
many elements of M to a single element of T.
59
Associated with each data set i is a collection of random variables
Q(i), which may be a vector. Let FQ (qjm) be the distribution of q under
model m. To establish identification on nonnegligible sets, one needs that, for
some true model m*,
Pr (jFQ qjm*  FQ qjmj > " > 0
for some " > 0 for all m 6 m*. This guarantees that there are observable differences between the data generating process for Q given m and for Q given m*. We
can also define this for FQ (qjt*) and FQ (qjt).

53

THE SCIENTIFIC MODEL OF CAUSALITY

FIGURE 1. Are elements in T uniquely determined from elements in I ?


Sometimes T M. Usually T consists of elements derived from M.

model is consistent with the data. By placing restrictions on models,


we can sometimes reduce the number of elements in Mh(i) if it has
multiple members. Let R M be a set of restricted models. It is
sometimes possible by imposing restrictions to reduce the number of
models consistent with the data. Recall that in the two-person model
of social interactions, if  12 0 and 21 0 we could uniquely
identify the remaining parameters under the other conditions maintained in Section 2.5. Thus R \ Mh(i) may contain only a single
element. Another way to solve this identification problem is to pick
another data source i0 2 I, which may produce more restrictions on
the class of admissible models. More information provides more
hoops for the model to jump through.
Going after a more limited class of objects such as features of a
model (t 2 T ) rather than the full model (m 2 M ) is another way to
secure unique identification. Let Mg(t) g1({t}) {m 2 M: g(m) t}.

54

HECKMAN

Necessary and sufficient conditions for the existence of a unique map


f : I T with the property f
h g are (a) h must map M onto I and (b)
for all i 2 I, there exists t 2 T such that Mh(i)  Mg(t). Condition (b)
means that even though one element i 2 I may be consistent with many
elements in M, so that Mh (i) consists of more than one element, it may
be that all elements in Mh(i) are mapped by g into a single element of
T. The map f is onto since g f
h and g is onto by assumption. In
order for the map f to be one-to-one, it is necessary and sufficient to
have equality of Mh(i) and Mg(t) instead of simply inclusion.
If we follow Marschaks maxim and focus on a smaller target
space T, it is possible that g maps the admissible models into a smaller
space. Thus the map f described above may produce a single element
even if there are multiple models m consistent with the data source i.
This would arise, for example, if for a given set of data i, we could
only estimate the mean 1 of Y1 up to a constant c and the mean 2 of
Y2 up to the same constant c. But we could uniquely identify the
element 1  2 2 T.60 In general, identifying elements of T is easier
than identifying elements of M. Thus, in Figure 1, even though many
models (elements of M) may be consistent with the same i 2 I, only
one element of T may be consistent with that i. I now turn to
empirical causal inference and illustrate the provisional nature of
causal inference.
4. THE PROVISIONAL NATURE OF CAUSAL INFERENCE61
This section develops the implicit assumptions underlying four widely
used methods of causal inference applied to data: (1) matching, (2)
control functions, (3) instrumental variable methods, and (4) the
method of directed acyclic graphs promoted by Pearl (2000) (or the
g-computation method of Robins 1989). It is not intented as an

60

Most modern analyses of identification assume that sample sizes are


infinite, so that enlarging the sample size is not informative. However, in any
applied problem this distinction is not helpful. Having a small sample (e.g. fewer
observations than regressors) can produce an identification problem. This definition combines task 3 and task 2 if we allow for samples to be finite.
61
Portions of this section are based on Heckman and Navarro (2004).

THE SCIENTIFIC MODEL OF CAUSALITY

55

exhaustive survey of the literature. I demonstrate the value of the


scientific approach to causality by showing how explicit analysis of
the choice of treatment (or the specification of the selection equations)
and the outcomes, including the relationship between the unobservables in the outcome and selection equations clarifies the implicit
assumptions being made in each method. This enables the analyst to
use behavioral theory aided by statistics to choose estimators and
interpret their output. This discussion also clarifies that each method
of inference makes implicit identifying assumptions in going from
samples to make inferences about models. There is no assumptionfree method of causal inference.62
I do not discuss randomization systematically except to note
that randomization does not in general identify distributions of treatment effects (Heckman 1992; Heckman and Smith 1998; Heckman,
Smith, and Clements 1997; Heckman and Vytlacil 2006b). Matching
implicitly assumes a randomization by nature in the unobservables
producing the choice treatment equation relative to the outcome equation, so my analysis of matching implicitly deals with randomization.
I focus primarily on identification of mean treatment effects in
this paper. Discussions of identification of distributions of treatment
effects are presented in Aakvik, Heckman, and Vytlacil (1999, 2005),
Carneiro, Hansen, and Heckman (2001, 2003), and Heckman and
Navarro (2006). I start by presenting a prototypical econometric
selection model.

4.1. A Prototypical Model of Treatment Choice and Outcomes


To focus the discussion, and to interpret the implicit assumptions underlying the different estimators presented in this paper, I present a benchmark model of treatment choice and treatment outcomes. For simplicity
I consider two potential outcomes (Y0, Y1). I drop the individual (!)
subscripts to avoid notational clutter. D 1 if Y1 is selected; D 0 if
Y0 is selected. Agents pick the realized outcome based on their evaluation of the outcomes, given their information. The agent picking the
treatment might be different from the person experiencing the outcome

62

This is true for experiments as well. See Heckman (1992).

56

HECKMAN

(e.g., the agent could be a parent choosing outcomes for the child). Let V
be the agents valuation of treatment. I write
V V W; UV

D 1V > 0;

20

where the W are factors (observed by the analyst) determining


choices, UV are the unobserved (by the analyst) factors determining
choice. Valuation function (20) is a centerpiece of the scientific model
of causality but is not specified in the statistical approach.
Potential outcomes are written in terms of observed variables
(X) and unobserved (by the analyst) outcome-specific variables
Y1 1 X; U1

21a

Y0 0 X; U0 :

21b

I assume throughout that U0, U1, and UV are continuous random


variables and that all means are finite.63 The individual level treatment effect is thus
 Y1  Y0 :
More familiar forms of (20), (21a), and (21b) are additively separable
expressions,
V V W UV EUV 0;

22a

Y1 1 X U1 EU1 0;

22b

Y0 0 X U0 EU0 0:

22c

Additive separability is not strictly required in modern econometric models


(e.g., see Matzkin 2003). However, I use the additively separable representation throughout most of this section because of its familiarity, noting
when it is a convenience and when it is an essential part of a method.
The distinction between X and Z is crucial to the validity of
many econometric procedures. In matching as conventionally
63

measure.

Strictly speaking, absolutely continuous with respect to the Lebesgue

THE SCIENTIFIC MODEL OF CAUSALITY

57

formulated there is no distinction between X and Z. The roles of X


and Z in alternative estimators are explored in this section.
A simple example will serve to fix ideas. It will enable me to
synthesize the main results of the first three sections of this paper and
lay the ground for this section.
Suppose that we use linear-in-parameters expressions. We write
the potential outcomes for the population as
Y1 X1 C1 U1

23a

Y0 X0 C0 U0 ;

23b

where we let X be the characteristics of persons and we let the 


depend on C1 and C0, the characteristics of the programs. These are
linear-in-parameters versions of equation (10) for s 0,1. The U1 and
U0 are the unobservables arising from omitted X, C1, and C0 components. Included among the X is 1 so that the characteristics of the
programs are allowed to enter directly and in interaction with the X.
By modeling how 1 and 0 depend on C1 and C0, we can answer
policy question P3 for new programs that offer new packages of C,
assuming we can account for the effects Ci on generating U1 and U0.
A version of the model most favorable to solving problems P2
and P3 writes
1 C1
C01
0 C0
C00 ;
where C1 and C0 are 1  J vectors of characteristics of programs, and
C01 and C00 are their transposes. Assuming that X is a 1  K vector of
person-specific characteristics,
is a K  J matrix. This specification
enables us to represent all of the coefficients of the outcome equations
in terms of a base set of generator characteristics.
For each fixed set of characteristics of a program, we can
model how outcomes are expected to differ when we change the
characteristics of the people participating in them (the X). This is an
ingredient for solving problem P3.
Equations (23a) and (23b) are in ex post all causes form.
For information set I , we can write the ex ante version as EY1 jI
and E(Y0 jI ) (see equation 11). The decision-making agent may be
uncertain about the X, the  i, the Ci, and the Ui. The ex ante version
reflects this uncertainty. Cunha, Heckman, and Navarro (2005a,b)

58

HECKMAN

provide examples of ex ante outcome models. Ex ante Marshallian


causal functions are defined in terms of variations in I . Ex post and ex
ante outcomes are connected by shock  (s, !), as in equation (12).
The choice equation may depend on expected rewards and
costs, as in Section 2.3. Let
V EY1  Y0  P1  P0 jI ;

24

where Pi is the price of participating in i and Pi Zi i. In the


special case of perfect foresight, I (U1 ; U0 ; C1 ; C0 ; X; Z;
; 1 ; 2 ).
To focus on some main ideas, suppose that we work with 1
and  0, leaving the Ci implicit. Substituting for the Pi in equation (24)
and for the outcomes (23a) and (23b), we obtain after some algebra
V EX1  0  Z1  0 U1 U0  1  0 j I ;
where I is the information set at the time the agent is making the
participation decision. Let W (X, Z), UW (U1  U0)  (1 0),
and (1   0,  (1  0)). We can then represent the choice
equation as
V EW UW jI ;
where
D 1V > 0:
Let UV be the random variable of UW conditional on I . For simplicity, we assume that agents know W (X, Z) but not all of the
components of UW when they make their treatment selection decisions. We also assume that the analyst knows W (X, Z).
The selection problem arises when D is correlated with (Y0, Y1).
This can happen if the observables or the unobservables in (Y0, Y1) are
correlated with or dependent on D. Thus there may be common
observed or unobserved factors connecting V and (Y0, Y1).
If D is not independent of (Y0, Y1), the observed (Y0, Y1) are
not randomly selected from the population distribution of (Y0, Y1). In
the Roy model, discussed in Section 1, 1 0 0, 1 0 0, and
selection is based on Y1 and Y0 (D 1(Y1 > Y0)). Thus we observe
Y1 if Y1 > Y0 and we observe Y0 if Y0  Y1.

THE SCIENTIFIC MODEL OF CAUSALITY

59

If conditioning on W makes (Y0, Y1) independent of D, selection on observables is said to characterize the selection process.64 This
is the motivation for the method of matching. If conditional on W,
(Y0, Y1) are not independent of D, then we have selection on unobservables and alternative methods must be used.
For the Roy model, Heckman and Honore (1990) show that
it is possible to identify the distribution of treatment outcomes
(Y1  Y0) under the conditions they specify. Randomization can
identify only the marginal distributions of Y0 and of Y1, not
the joint distribution of (Y1  Y0) or the quantiles of (Y1  Y0).
Thus, under its assumptions, the Roy model is more powerful
than randomization in producing the distributional counterfactuals.65
The role of the choice equation is to motivate and justify the
choice of an evaluation estimator. This is a central feature of
the econometric approach that is missing from the statistical and
epidemiological literature on treatment effects. Heckman and
Smith (1998), Aakvik, Heckman, and Vytlacil (2005), Carneiro,
Hansen, and Heckman (2003), and Cunha, Heckman, and Navarro
(2005a,b) extend these results to estimate distributions of treatment
effects.

4.2. Parameters of Interest


There are many different treatment parameters that can be derived
from this model if U1 6 U0 and agents know or partially anticipate
U0, U1 in making their decisions (Heckman and Robb 1985; Heckman
1992; Heckman, Smith, and Clements 1997: Heckman 2001; Heckman
and Vytlacil 2000; Cunha, Heckman, and Navarro 2005a,b). For
specificity, I focus on certain means because they are traditional. As
noted in Section 2 and in Heckman and Vytlacil (2000, 2005) and
Heckman (2001), the traditional means often do not answer interesting social and economic questions.

64

See Heckman and Robb (1985).


The same analysis applies to matching, which cannot identify the
distributions of (Y1  Y0) or derived quantiles.
65

60

HECKMAN

The traditional means conditional on covariates are as follows:


Average Treatment Effect ATE : EY1  Y0 jX
Treatment on the Treated TT : EY1  Y0 jX; D 1
Marginal Treatment Effect MTE : EY1  Y0 jX; Z; V 0:
The MTE is the marginal treatment effect introduced into the
evaluation literature by Bjorklund and Moffitt (1987). It is the average
gain to persons who are indifferent to participating in sector 1 or sector
0 given X, Z. These are persons at the margin, defined by (W) so Z
plays a role in the definition of the parameter by fixing V(W) in
equation (20) or equation (22a) and hence fixing UV. It is a version
of EOTM as defined in Section 1. An alternative definition in this setup
is MTE E(Y1  Y0jX, UV). Heckman and Vytlacil (1999, 2005,
2006b) show how the MTE can be used to construct all mean treatment
parameters, including the policy relevant treatment parameters, under
the conditions specified in their papers. These parameters can be
defined for the population as a whole not conditioning on X or Z.66

4.3. The Selection Problem Stated in Terms of Means


Let Y DY1 (1  D)Y0. Samples generated by choices have the
following means which are assumed to be known:
EYjX; Z; D 1 EY1 jX; Z; D 1
and
EYjX; Z; D 0 EY0 jX; Z; D 0
for outcomes Y1 for participants and the outcomes Y0 for nonparticipants,
respectively. In addition, choices are observed so that in large samples
Pr(D 1jX, Z) is knownthat is, the probability of choosing treatment is
known. From the sample data, we can also construct
EY1 jX; D 1 and
66

EY0 jX; D 0:

The average marginal


treatment effect is
Z
EY1  Y0 jX; Z; V 0fX; ZjV 0dXdZ:

EY1  Y0 jV 0

THE SCIENTIFIC MODEL OF CAUSALITY

61

The conditional biases from using the difference of these means


to construct the three parameters studied in this paper are
Bias TT EYjX; D 1  EYjX; D 0  EY1  Y0 jX; D 1
EY0 jX; D 1  EY0 jX; D 0:
In the case of additive separability
Bias TT EU1 jX; D 1  EU0 jX; D 0:
For ATE,
Bias ATE EYjX; D 1  EYjX; D 0  EY1  Y0 jX:
In the case of additive separability
Bias ATE EU1 jX; D 1  EU1 jXEU0 jX; D 0  EU0 jX:

For MTE,
Bias MTE EYjX; Z; D 1  EYjX;Z; D 0
 EY1  Y0 jX;Z;V 0
EU1 jX; Z;D 1  EU1 jX; Z;V 0
 EU0 jX; Z; D 0  EU0 jX; Z; V 0;
for the case of additive separability in outcomes. The MTE is defined for
a subset of persons indifferent between the two sectors and so is defined
for X and Z. The bias is the difference between average U1 for participants and marginal U1 minus the difference between average U0 for
nonparticipants and marginal U0. Each of these terms is a bias that can
be called a selection bias. These biases can be defined conditional on X
(or X and Z or X, Z, and V in case of the MTE) or unconditionally.

4.4. How Different Methods Eliminate the Bias


In this section I consider the identification conditions that underlie
matching, control functions, and instrumental variable methods to

62

HECKMAN

identify the three parameters using the data on mean outcomes. I also
briefly discuss the method of directed acyclic graphs or the g-computation method for one causal parameter. I discuss sources of unobservables, implicit assumptions about how unobservables are eliminated as
sources of selection problems, and the assumed relationship between
outcomes and choice equations. I start with the method of matching.

4.4.1. Matching
The method of matching as conventionally formulated makes
no distinction between X and Z. Define the conditioning set as
W (X, Z). The strong form of matching advocated by Rosenbaum
and Rubin (1983) and in numerous predecessor papers, assumes that
Y1 ; Y0 ?? DjW

M-1

0 < PrD 1jW PW < 1;

M-2

and

where ?? denotes independence given the conditioning variables after


j. P(W) is the probability of selection into treatment and is sometimes
called the propensity score. Condition (M-2) implies that the mean
treatment parameters can be defined for all values of W (i.e., for each
W, in very large samples, there are observations for which we observe a
Y0 and other observations for which we observe a Y1). Rosenbaum and
Rubin (1983) show that under (M-1) and (M-2)
Y1 ; Y0 ?? DjPW:

M-3

This reduces the dimensionality of the matching problem. They assume


that P is known. When it is not known, it is necessary to estimate it.
Nonparametric estimation of P(W) restores the dimensionality
problem but shifts it to the estimation of P(W).67 Under these
67
Rosenbaum (1987) or Rubin and Thomas (1992) consider the distribution of the matching estimator when P is estimated under special assumptions
about the distribution of the matching variables. Papers that account for estimated P under general conditions include Heckman, Ichimura, and Todd (1997,
1998) and Hahn (1998).

THE SCIENTIFIC MODEL OF CAUSALITY

63

assumptions, conditioning on P eliminates all three biases defined in


Section 4.3 for parameters defined conditional on P because
EY1 jD 0; PW EY1 jD 1; PW EY1 jPW
EY0 jD 1; PW EY0 jD 0; PW EY0 jPW:
Thus for TT one can identify counterfactual mean E(Y0jD 1, P(W))
from E(Y0jD 0, P(W)). In fact, one only needs the weaker condition
Y0??DjP(W) to remove the bias68 because E(Y1jD 1, P(W)) is
known, and only E(Y0jD 1,P(W)) is unknown. From the observed
conditional means one can form ATE. Since the conditioning is on
P(W), the parameter is defined conditional on it and not X or (X, Z).
Integrating out P(W) produces unconditional ATE. Integrating out
P(W) given D 1 produces unconditional TT.69
Observe that since ATE TT for all X, Z under (M-1) and
(M-2), the effect for the average person participating in a program is
the same as the effect for the marginal person, conditional on W, and
there is no bias in estimating MTE.70 The strong implicit assumption
that the marginal participant in a program gets the same return as the
average participant in the program, conditional on W, is an unattractive implication of these assumptions (see Heckman 2001 and
Heckman and Vytlacil 2005, 2006a,b). The method assumes that all
of the dependence between UV and (U1, U0) is eliminated by conditioning on W,
UV ?
? U1; U0 jW:
This motivates the term selection on observables introduced in
Heckman and Robb (1985, 1986).
Assumption (M-2) has the unattractive feature that if the analyst has too much information about the decision of who takes treatment so that P(W) 1 or 0, the method breaks down because people
cannot be compared at a common W. The method of matching
68

See Heckman, Ichimura, and Todd (1997) and Abadie (2003).


To estimate the parameters conditional on W, one cannot use P(W)
but must use the full W vector.
70
As demonstrated in Carneiro (2002), one can still distinguish marginal
and average effects in terms of observables.
69

64

HECKMAN

assumes that, given W, some unspecified randomization device allocates people to treatment. The fact that the cases P(W) 1 and
P(W) 0 must be eliminated suggests that methods for choosing
which variables enter W based on the fit of the model to data on
choices (D) are potentially problematic; see Heckman and Navarro
(2004) and Heckman and Vytlacil (2005) for further discussion of this
point.
What justifies (M-1) or (M-3)? Absent an explicit theoretical
model of treatment assignment and an explicit model of the sources of
randomness, analysts are unable to justify the assumption except by
appeal to convenience. Because there are no exclusion restrictions in
the observables, the only possible source of variation in D given W are
the unobservable elements generating D. These elements are assumed
to act like an ideal randomization that assigns person to treatment but
is independent of (U1, U2), the unobservables generating (Y0, Y1),
given W.
If agents partially anticipate the benefits of treatment and
make enrollment decisions based on these anticipations, (M-1) or
(M-3) is false. In the extreme case of the Roy model, where
D 1(Y1 > Y0), (M-1) or (M-3) is certainly false. Even if agents are
only imperfectly prescient but can partially forecast (Y1,Y0) and use
that information in deciding whether or not to participate, (M-1) or
(M-3) is false.
Without a model of interventions justifying these assumptions,
and without a model of the sources of unobservables, (M-1) or (M-3)
cannot be justified. The model cannot be tested without richer sources
of data.71 Judgments about whether agents are as ignorant about
potential outcomes given W, as is assumed in (M-1) or (M-3), can
only be settled by the theory unless it is possible to randomize persons
into treatment, and randomization does not change the outcome
that is, under assumption (A-2). The matching model makes strong
implicit assumptions about the unobservables.
In the recent literature, the claim is sometimes made that matching is for free (e.g., see Gill and Robins 2001). The idea underlying
this claim is that since E (Y0jD 1, W) is not observed, we might as
well set it to E (Y0jD 0, W), an implication of (M-1). This argument
71

See Heckman, Ichimura, Smith, and Todd (1998) for a test of matching assumptions using data from randomized trials.

THE SCIENTIFIC MODEL OF CAUSALITY

65

is correct so far as data description goes. Matching imposes just-identifying restrictions and in this senseat a purely empirical levelis as
good as any other just-identifying assumption in describing the data.
However, the implied behavioral restrictions are not for free.
Imposing thatconditional on X and Z or conditional on P(W) the
marginal person entering a program is the same as the average person
is a strong and restrictive implication of the conditional independence
assumptions and is not a for free assumption in terms of its behavioral
content.72 In the context of estimating the economic returns to schooling, it implies that, conditional on W, the economic return to schooling
for persons who are just at the margin of going to school are the same as
the return for persons with strong preferences for schooling.
Introducing a distinction between X and Z allows the analyst
to overcome the problem arising from perfect prediction of treatment
assignment for some values of (X, Z) if there are some variables Z not
in X. If P is a nontrivial function of Z (so P(X, Z) varies with Z for all
X) and Z can be varied independently of X for all points of support of
X,73 and if outcomes are defined solely in terms of X, the problem of
perfect classification can be solved. Treatment parameters can be
defined for all support values of X since for any value (X, Z) that
perfectly classifies D, there is another value (X, Z0 ), Z0 6 Z, that does
not (see Heckman, Ichimura, and Todd 1997).
Offsetting the disadvantages of matching, the method of
matching with a known conditioning set that satisfies (M-1) does
not require separability of outcome or choice equations into observable and unobservable components, exogeneity of conditioning variables, exclusion restrictions, or adoption of specific functional forms
of outcome equations. Such assumptions are commonly used in conventional selection (control function) methods and conventional
applications of IV although recent work in semiparametric estimation

72
As noted by Heckman, Ichimura, Smith, and Todd (1998), if one seeks
to identify E (Y1  Y0jD 1, W) one only needs to impose a weaker condition
[E (Y0jD 1, W) E(Y0jD 0, W)] or Y0 ?? DjW rather than (M-1). This
imposes the assumption of no selection on levels of Y0 (given W) and not the
assumption of no selection on levels of Y1 or on Y1  Y0, as (M-1) does.
Marginal can be different from average in this case.
73
A precise sufficient condition is that Supp (ZjX) Supp (Z). We can
get by with a weaker condition that in any neighborhood of X, there is a Z* such
that 0 < Pr(D 1jX, Z*) < 1, and that Z* is in the support of ZjX.

66

HECKMAN

relaxes many of these assumptions, as I note below (see also Heckman


and Vytlacil 2005, 2006b). Moreover, the method of matching does
not strictly require (M-1). One can get by with weaker mean independence assumptions,
EY1 jW; D 1 EY1 jW;
EY0 jW; D 0 EY0 jW;

M-10

in the place of the stronger (M-1) conditions. However, if (M-10 ) is


invoked, the assumption that one can replace W by P (W) does not
follow from the analysis of Rosenbaum and Rubin, and is an additional new assumption.

4.4.2. Control Functions


The principle motivating the conventional method of control functions is different. (See Heckman 1976, 1978, 1980 and Heckman
and Robb 1985, 1986, where this principle was first developed.)
Like matching, it works with conditional expectations of (Y1,Y0)
given (X, Z and D). Conventional applications of the control
function method assume additive separability that is not required
in matching. Strictly speaking, additive separability in the outcome
equation is not required in the application of control functions
either.74 What is required is a model relating the outcome unobservables to the observables, including the choice of treatment. The
method of matching assumes that, conditional on the observables
(X, Z), the unobservables are independent of D.75 For the additively separable case, control functions based on the principle of
modeling the conditional expectations of Y1 and Y0 given X, Z,
and D can be written as
EY1 jX; Z; D 1 1 X EU1 jX; Z; D 1
EY0 jX; Z; D 0 0 X EU0 jX; Z; D 0:

74
Examples of nonseparable selection models are found in Cameron and
Heckman (1998).
75
Or mean independent in the case of mean parameters.

67

THE SCIENTIFIC MODEL OF CAUSALITY

In the method of control functions if one can model


E(U1jX, Z, D 1) and E(U0jX, Z, D 0) and these functions
can be independently varied against 1 (X) and 0 (X) respectively,
one can identify 1 (X) and 0 (X) up to constant terms.76 Nothing
in the method intrinsically requires that X or Z be stochastically
independent of U1 or U0, although conventional methods often
assume this.
If one assumes that (U1, UV)??(X, Z) and adopts equation
(22a) as the treatment choice model augmented so X and Z are
determinants of treatment choice, one obtains
EU1 jX; Z; D 1 EU1 jUV  V X; Z K1 PX; Z;
so the control function depends only on P(X, Z). By similar reasoning, if (U0, UV) ?? (X, Z),
EU0 jX; Z; D 0 EU0 jUV < V X; Z K0 PX; Z
and the control function depends only on the probability of selection
(the propensity score). The key assumption needed to represent the
control function solely as a function of P(X, Z) is
U1 ; U0 ; UV ?? X; Z:

C-1

Under this condition


EY1 jX; Z; D 1 1 X K1 PX; Z
EY0 jX; Z; D 0 0 X K0 PX; Z

76
Heckman and Robb (1985, 1986) introduce this general formulation of control functions. The identifiability requires that the members of the
pairs (1(X), E (U1jX, Z, D 1)) and (0(X), E (U0jX, Z, D 0)) be variation free so that they can be independently varied against each other; see
Heckman and Vytlacil (2006a, b) for a precise statement of these conditions.

68

HECKMAN

with limP!1 K1 (P) 0 and limP!0 K0 (P) 0 where it is assumed


that Z can be independently varied for all X, and the limits are
obtained by changing Z while holding X fixed.77 These limit results
simply state that when the values of X,Z are such that the probability of being in a sample is 1, there is no selection bias. One can
approximate the K1(P) and K0(P) terms by polynomials in P
(Heckman 1980; Heckman and Robb 1985, 1986; Heckman and
Hotz 1989).
If K1(P(X, Z)) can be independently varied from 1(X) and
K0(P(X, Z)) can be independently varied from 0(X), one can
identify 1(X) and 0(X) up to constants. If there are limit sets
Z0 and Z1 such that for each X limZ!Z0 P(X; Z) 0 and
limZ!Z1 P(X; Z) 1, then one can identify these constants, since in
those limit sets we identify 1 (X) and 0 (X).78 Under these conditions, it is possible to nonparametrically identify all three conditional treatment parameters:
ATEX 1 X  0 X
TTX; D 1 1 X  0 X EU1  U0 jX; D 1
1 X  0 X EZjX;D1




1P
K0 PX; Z ;79
K1 PX; Z
P

77

More precisely, assume that Supp (ZjX) Supp (Z) and that limit
sets of Z, Z0, and Z1 exist such that as Z Z0, P(Z, X) 0 and as Z Z1,
P(Z, X) 1. This is also the support condition used in the generalization of
matching by Heckman, Ichimura, and Todd (1997).
78
This condition is sometimes called identification at infinity; see
Heckman (1990) or Andrews and Schafgans (1998).
79
Since
EU0 0
EU0 jD 1; X; ZPX; Z EU0 jD 0; X; Z1  PX; Z
EU0 jD 1; X; Z 

1  PX; Z
1  PX; Z
EU0 jD 0; X; Z 
K0 PX; Z
PX; Z
PX; Z

See Heckman and Robb (1986). The expression EZjX, D 1 integrates out Z for a
given X, D 1.

THE SCIENTIFIC MODEL OF CAUSALITY

69

MTEX; Z; V 0 1 X  1 X EU1  U0 j V Z; X
UV
1 X  0 X
@ EU1  U0 jX; Z; D 1PX; Z 80
:

@PX; Z
Unlike the method of matching, the method of control functions
allows the marginal treatment effect to be different from the average
treatment effect or from the effect of treatment on the treated (i.e.,
the second term on the right-hand side of the first equation for
MTE(X, Z, U 0) is, in general, nonzero). Although conventional
practice is to derive the functional forms of K0(P) and K1(P) by making
distributional assumptions (e.g., normality or other conventional
distributional assumptions about (U0, U1, UV); see Heckman, Tobias,
and Vytlacil 2001, 2003), this is not an intrinsic feature of the method
and there are many non-normal and semiparametric versions of this
method (see Powell 1994 or Heckman and Vytlacil 2006a,b for surveys).
Without invoking parametric assumptions, the method of control functions requires an exclusion restriction (a Z not in X) to
achieve nonparametric identification.81 Without any functional form
assumptions, one cannot rule out a worst-case analysis wherefor
example, if X Z, K1(P(X)) a(X) where a is a scalar. Then, there

80
As established in Heckman and Vytlacil (2000, 2005) and Heckman
(2001), under assumption (C-1) and additional regularity conditions

Z
EU1 U0 jX;Z;D1PX;Z

PX;Z 1



U1 U0 f U1U0 jU*V dU1 U0 dU*V ;

where U*VFV (UV ), so



@EU1 U0 jX;Z;D1PX;Z 
E U1 U0 jU*VPX;Z :
@PX;Z
The third expression follows from algebraic manipulation. Expressions conditional on X and V 0 are obtained by integrating out Z conditional on X and
V 0.
81
For many common functional forms for the distributions of unobservables, no exclusion is required.

70

HECKMAN

is perfect collinearity between the control function and the conditional


mean of the outcome equation, and it is impossible to control for
selection with this method. Even though this case is not generic, it is
possible. The method of matching does not require an exclusion
restriction because it makes a stronger assumption, which we clarify
below. Without additional assumptions, the method of control functions requires that, for some Z values for each X, P(X, Z) 1 and
P(X, Z) 0 to achieve full nonparametric identification.82 The conventional method of matching excludes this case.
Both methods require that treatment parameters be defined on
a common support that is the intersection of the supports of X given
D 1 and X given D 0:
Supp XjD 1 \ Supp XjD 0:
A similar requirement is imposed on the generalization of matching
with exclusion restrictions introduced in Heckman, Ichimura,
Smith, and Todd (1998). Recall that exclusion (adding a Z in the
probability of treatment equation that is not in the outcome
equation where Pr(D 1jX, Z) is the choice probability), both
in matching and selection models, enlarges the set of X values
that satisfy this condition. If P(X, Z) depends on Z, then even if
P(X, Z) 1 for some Z z it can be that P(X, Z) < 1 for Z z0
if z 6 z0 . A similar argument applies to P(X, Z) 0 for Z z00 but
P(X, Z) > 0 for Z z000 if z00 6 z000 . This requires the existence of
such Z values in the neighborhood of all values of X, Z such that
P(X, Z) 0 or 1.
In the method of control functions, P(X, Z) is a conditioning
variable used to predict U1 conditional on D, X, and Z and U0
conditional on D, X, and Z. In the method of matching, it is used
to characterize the stochastic independence between (U0, U1) and D.
In the method of control functions, as conventionally applied, (U0, U1)
?? (X, Z), but this assumption is not intrinsic to the method.83

82
Symmetry of the errors can be used in place of the appeal to limit sets
that put P(X, Z) 0 or P(X, Z) 1; see Chen (1999).
83
Relaxing it, however, requires that the analyst model the dependence
of the unobservables on the observables and that certain variation-free conditions
are satisfied; see Heckman and Robb (1985).

THE SCIENTIFIC MODEL OF CAUSALITY

71

This assumption plays no role in matching if the correct conditioning


set is known (i.e., one that satisfies (M-1) and (M-2)). However, as
noted in Heckman and Navarro (2004), exogeneity plays a key role in
devising rules to select appropriate conditioning variables. The method
of control functions does not require that (U0, U1) ?? Dj(X, Z), which
is a central requirement of matching. Equivalently, the method of
control functions does not require
U0 ; U1 ?? UV jX; Z
whereas matching does. Thus matching assumes access to a richer set
of conditioning variables than is assumed in the method of control
functions.
The method of control functions is more robust than the
method of matching, in the sense that it allows for outcome unobservables to be dependent on D even after conditioning on (X, Z), and it
models this dependence, whereas the method of matching assumes no
such dependence. Matching under the assumed conditions is a special
case of the method of control functions84 in which under assumptions
(M-1) and (M-2),
EU1 jX; Z; D 1 EU1 jX; Z
EU0 jX; Z; D 0 EU0 jX; Z:
In the method of control functions in the case when (X, Z) ?? (U0, U1, UV)
EYjX; Z; D EY1 jX; Z; D 1D EY0 jX; Z; D 01  D
0 X 1 X  0 XD
EU1 jX; Z; D 1D EU0 jPX; Z; D 01  D
0 X 1 X  0 XD
EU1 jPX; Z; D 1D EU0 jPX; Z; D 01  D
0 X 1 X  0 X K1 PX; Z  K0 PX; ZD
K0 PX; Z:

84

See Aakvik et al. (2005); Carneiro et al. (2003); and Cunha et al.
(2005a, 2005b) for a generalization of matching that allows for selection on
unobservables by imposing a factor structure on the errors and estimating the
distribution of the unobserved factors.

72

HECKMAN

To identify 1(X)  0(X), the average treatment effect, one must


isolate it from K1(P(X, Z)) and K0(P(X, Z)). The coefficient on D in
this regression does not correspond to any one of the treatment effects
presented above.
Under assumptions (M-1) and (M-2) of the method of matching, one may write expressions conditional on P(W):
EYjPW; D 0 PW

1 PW  0 PW EU1 jPW  EU0 jPW


D fEU0 jPWg:

Notice that if the analyst further invokes (C-1)


EYjPW; D 0 PW 1 PW  0 PWD;
since E(U1jP(W)) E(U0jP(W)) 0. A parallel argument can be
made conditioning on X and Z instead of P(W).
Under the assumptions that justify matching, treatment
effects ATE or TT (conditional on P(W)) are identified from the
coefficient on D in either of the two preceding equations. It is not
necessary to invoke (C-1) in the application of matching although
it simplifies expressions. One can define the parameters conditional
on X, allowing the X to be endogenous. Condition (M-2) guarantees that D is not perfectly predictable by W so the variation in D
identifies the treatment parameter. Thus the coefficient on D in the
regression associated with the more general control function model
does not correspond to any treatment parameter whereas the coefficient on D in the regression associated with matching corresponds to a treatment parameter under the assumptions of the
matching model. Under (C-1), 1(P(W))  0(P(W)) ATE and
ATE TT MTE, so the method of matching identifies all of
the (conditional on P(W)) mean treatment parameters.85 Under the
assumptions justifying matching, when means of Y1 and Y0 are the

85
This result also holds if (C-1) is not satisfied, but then the treatment
effects include
EU1 jPW  EU0 jPW
.

73

THE SCIENTIFIC MODEL OF CAUSALITY

parameters of interest, and W satisfies (M-1) and (M-2), the bias


terms defined in Section 4.3 vanish. They do not in the more
general case considered in the method of control functions. The
vanishing of the bias terms in matching is the mathematical counterpart of the randomization implicit in matching: conditional on
W or P(W), (U1, U0) are random with respect to D. The method of
control functions allows them to be nonrandom with respect to D.
In the absence of functional form assumptions, an exclusion
restriction is required in the analysis of control functions to separate out K0(P(X, Z)) from the coefficient on D. Matching produces
identification without exclusion restrictions whereas identification
with exclusion restrictions is a central feature of the control function method in the absence of functional form assumptions. The
implicit randomization in matching plays the role of an exclusion
restriction in the method of instrumental variables.
The work of Rosenbaum (1995) and Robins (1997) implicitly
recognizes that the control function approach is more general than the
matching approach. Their sensitivity analyses for matching when
there are unobserved conditioning variables are, in their essence,
sensitivity analyses using control functions.86 Aakvik, Heckman,
and Vytlacil (2005), Carneiro, Hansen, and Heckman (2003), and
Cunha, Heckman, and Navarro (2005a) explicitly model the relationship between matching and selection models using factor structure
models, treating the omitted conditioning variables as unobserved
factors and estimating their distribution.
Tables 2 and 3 perform sensitivity analyses under different
assumptions about the parameters of the underlying selection
model. In particular, I assume that the data are generated by the
model of equations (22a)(22c), with (22c) having the explicit representation
V Z UV ;
U1 ; U0 ; UV 0 N0; 
corr Uj ; UV
jV
var Uj 2j ;
86

j f0; 1g:

See also Vijverberg (1993), who performs a sensitivity analysis in a


parametric selection model with an unidentified parameter.

74

HECKMAN

TABLE 2
Mean Bias for Treatment on the Treated

0V

Average Bias ( 0 1)

Average Bias ( 0 2)

1.7920
1.3440
0.8960
0.4480
0.0000
0.4480
0.8960
1.3440
1.7920

3.5839
2.6879
1.7920
0.8960
0.0000
0.8960
1.7920
2.6879
3.5839

1.00
0.75
0.50
0.25
0.00
0.25
0.50
0.75
1.00

BIASTT
0V * 0 * M(p)
M(p)

(1 (p))
p*(1  p)

I assume no X and that Z ?? (U1, U0, UV). Using the formulas


presented in the appendix of Heckman and Navarro (2004), one can
write the biases conditional on Z z as
Bias TTZ z Bias TTPZ pz 0
0V Mpz
Bias ATEZ z Bias ATEPZ pz
Mpz 1
1V 1  pz 0
0V pz
Bias MTEZ z Bias MTEPZ pz
Mpz 1
1V 1  pz 0
0V pz
 1 1  pz 1
1V  0
0V 
1

(1p(z)))
where M(p(z)) (
p(z)(1p(z)) , () and () are the probability density
function (pdf) and cumulative distribution function (cdf) of a standard
normal random variable and p(z) is the propensity score evaluated at
Z z. I assume that 1 0 so that the true average treatment effect
is zero.
I simulate the mean bias for TT (Table 2) and ATE (Table
3) for different values of the
jV and j. The results in the tables
show that, as one lets the variances of the outcome equations
grow, the value of the mean bias that one obtains can become
substantial. With larger correlations come larger biases. These

1.00

1.7920
1.5680
1.3440
1.1200
0.8960
0.6720
0.4480
0.2240
0

2.6879
2.4639
2.2399
2.0159
1.7920
1.5680
1.3440
1.1200

0V

1.00
0.75
0.50
0.25
0
0.25
0.50
0.75
1.00

1.00
0.75
0.50
0.25
0
0.25
0.50
0.75

2.2399
2.0159
1.7920
1.5680
1.3440
1.1200
0.8960
0.6720

1.5680
1.3440
1.1200
0.8960
0.6720
0.4480
0.2240
0
0.2240

0.75

1.7920
1.5680
1.3440
1.1200
0.8960
0.6720
0.4480
0.2240

1.3440
1.1200
0.8960
0.6720
0.4480
0.2240
0
0.2240
0.4480

0.50

1.3440
1.1200
0.8960
0.6720
0.4480
0.2240
0
0.2240

0.8960
0.6720
0.4480
0.2240
0
0.2240
0.4480
0.6720
0.8960

0.8960
0.6720
0.4480
0.2240
0
0.2240
0.4480
0.6720

1V ( 1 2)

1.1200
0.8960
0.6720
0.4480
0.2240
0
0.2240
0.4480
0.6720

1V ( 1 1)

0.25

( 0 1)

0.4480
0.2240
0
0.2240
0.4480
0.6720
0.8960
1.1200

0.6720
0.4480
0.2240
0
0.2240
0.4480
0.6720
0.8960
1.1200

0.25

TABLE 3
Mean Bias for Average Treatment Effect

0
0.2240
0.4480
0.6720
0.8960
1.1200
1.3440
1.5680

0.4480
0.2240
0
0.2240
0.4480
0.6720
0.8960
1.1200
1.3440

0.50

0.4480
0.6720
0.8960
1.1200
1.3440
1.5680
1.7920
2.0159

0.2240
0
0.2240
0.4480
0.6720
0.8960
1.1200
1.3440
1.5680

0.75

0.8960
1.1200
1.3440
1.5680
1.7920
2.0159
2.2399
2.4639

0
0.2240
0.4480
0.6720
0.8960
1.1200
1.3440
1.5680
1.7920

continued

1.00

THE SCIENTIFIC MODEL OF CAUSALITY

75

76

HECKMAN

tables demonstrate the greater generality of the control function


approach. Even if the correlation between the observables and the
unobservables (
jV) is small, so that one might think that selection
on unobservables is relatively unimportant, one still obtains substantial biases if one does not control for relevant omitted conditioning variables. Only for special values of the parameters can
one avoid bias by matching. These examples also demonstrate that
sensitivity analyses can be conducted for analysis based on control
function methods even when they are not fully identified, as noted
by Vijverberg (1993).

4.4.3. Instrumental Variables


Both the method of matching and the method of control functions
work with E(YjX, Z, D) and Pr(D 1jX, Z). The method
of instrumental variables works with E(YjX, Z) and Pr(D
1jX, Z). There are two versions of the method of instrumental
variables: (1) conventional linear instrumental variables and (2)
local instrumental variables (LIV) (Heckman and Vytlacil 1999, 2000,
2006b; Heckman 2001). LIV is equivalent to a semiparametric selection
model (Vytlacil 2002; Heckman and Vytlacil 2005, 2006b). It is an alternative way to implement the principle of control functions. LATE
(Imbens and Angrist 1994) is a special case of LIV under the conditions
I specify below.
I first consider the conventional method of instrumental variables. In this framework, P(X, Z) arises less naturally than it does in
the matching and control function approaches. Z is the instrument
and P(X, Z) is a function of the instrument.
Using the model of equations (22b) and (22c), I obtain
Y DY1 1  DY0
0 X 1 X  0 X U1  U0 D U0
0 X XD U0 ;
where (X) 1(X)  0(X) U1  U0. When U1 U0, we obtain
the conventional model to which IV is typically applied with

THE SCIENTIFIC MODEL OF CAUSALITY

77

D correlated with U0. Standard instrumental variable conditions


apply and P(X,Z) is a valid instrument if
EU0 jPX; Z; X EU0 jX87

IV-1

PrD 1jX; Z

IV-2

and

is a nontrivial function of Z for each X. When U1 6 U0 but D ??


(U1 U0)jX (or alternatively UV ?? (U1  U0)jX), then the same two
conditions identify (conditional on X):
ATE X EY1  Y0 jX EXjX
TT X EY1  Y0 jX; D 1 EY1  Y0 jX EX j X
MTEX
and the marginal equals the average conditional on X and Z. The
requirement that D ?? (U1  U0)jX is strong and assumes that agents
do not participate in the program on the basis of any information
about unobservables in gross gains (Heckman and Robb 1985, 1986;
Heckman 1997).88
How reasonable are the identifying assumptions of IV? An
appeal to behavioral theory helps. Consider the use of draft lottery
numbers as instruments (Z) for military service (Z 1 if served in the
army; Z 0 otherwise). The question is how does military service
affect earnings? (Angrist 1991). If agents participate in the military
87
Observe that it is not required that E (U0jX) 0. We can write the IV
estimator in the population as

EYjPX x; Z z pz ; X x  EYjPX x; Z z0 pz0 ; X x


PX x; Z z  PX x; Z z0
0 X XPX x; Z z EU0 jX  0 X XPX x; Z z  EU0 jX

PX x; Z z  PX x; Z z0
x

IV x

Thus it is not necessary to assume that E (U0 j X) 0.


88

We define ATE conditional on X as


EY1  Y0 jX x 1 X  0 X EU1  U0 jX x:

78

HECKMAN

based in part on the gain in the outcome measure (Y1,Y0) (e.g., the
difference in earnings) and this is a nondegenerate random variable,
then (IV-1) is violated and IV does not identify ATE. The validity of
the estimator is conditional on an untestable behavioral assumption.
Similar remarks apply to LATE as developed by Imbens and Angrist
(1994) and popularized by Angrist, Imbens, and Rubin (1996); see
Heckman and Vytlacil (1999, 2000, 2005), and Vytlacil (2002) for
more discussion of the implicit behavioral assumptions underlying
LATE.
The more interesting case for many problems arises when
U1 6 U0 and D (U1  U0) so agents participate in a program based
at least in part on factors not measured by the economist. To identify
ATE(X) using IV, it is required that
EU0 DU1  U0 jPX; Z; X EU0 DU1  U0 jX

IV-3

and condition (IV-2) (Heckman and Robb 1985, 1986; Heckman


1997). To identify TT(X) using IV, it is required that
EU0 DU1  U0  EU0 DU1  U0 jXjPX; Z; X
EU0 DU1  U0  EU0 DU1  U0 jXjX

IV-4

and condition (IV-2). No simple conditions exist to identify the MTE


using linear instrumental variables methods in the general case where
D (U1  U0)jX, Z. Heckman and Vytlacil (2001, 2005, 2006a,b)
characterize what conventional IV estimates in terms of a weighted
average of MTEs.
The conditions required to identify ATE using P as an instrument may be written in the following alternative form:
EU0 jPX; Z; X EU1  U0 jD 1; PX; Z; XPX; Z
EU0 jX EU1  U0 jD 1; XPX; Z:
If U1 U0 (everyone with the same X responds to treatment in
the same way) or (U1  U0) ?? DjP(X, Z), X (people do not participate
in treatment on the basis of unobserved gains), then these conditions
are the standard instrumental variable conditions. In general, the
conditions are not satisfied by economic choice models, except under

THE SCIENTIFIC MODEL OF CAUSALITY

79

special cancellations. If Z is a determinant of choices, and U1  U0


is in the agents choice set (or is only partly correlated with information in the agents choice set), then this condition is not satisfied
generically.
These identification conditions are fundamentally different
from the conditions required to justify matching and control function
methods. In matching, the essential condition for means (conditioning
on X and P(X, Z)) is
EU0 jX; D 0; PX; Z EU0 jX; PX; Z
and
EU1 jX; D 1; PX; Z EU1 jX; PX; Z:
These conditions require that, conditional on P(X, Z) and X, U1, and U0
are mean independent of UV (or D). If (C-1) is invoked, 1(W) and
0(W) are the conditional means of Y1 and Y0 respectively, the two
preceding expressions are zero. However, as I have stressed repeatedly,
(C-1) is not strictly required in matching.
The method of control functions models and estimates
the dependence of U0 and U1 on D rather than assuming that
it vanishes like the method of matching. The method of linear
instrumental variables requires that the composite error term
U0 D(U1  U0) be mean independent of Z (or P(X, Z)), given X.
Essentially, these conditions require that the dependence of U0 and
D(U1  U0) on Z vanish through conditioning on X. Matching requires
that U1 and U0 are independent of D given (X, Z). These conditions are
logically distinct. One set of conditions does not imply the other
set (Heckman and Vytlacil 2006a,b). They are justified by different
a priori assumptions. Hence the provisional nature of causal knowledge.
Assuming finite means, local instrumental variables methods developed by Heckman and Vytlacil (1999, 2001, 2005) estimate all three
treatment parameters in the general case where (U1  U0) 6?? Dj(X, Z)
under the following additional conditions
D Z

is a non-degenerate random variable given X


exclusion restriction

LIV-1

80

HECKMAN

U0 ; U1 ; UV ?? ZjX

LIV-2

0 < PrD 1jX < 1

LIV-3

Supp PD 1jX; Z 0; 1:

LIV-4

Under these conditions


@EYjX; PX; Z
MTEX; PX; Z; V 0:89
@PX; Z
Only (LIV-1)(LIV-3) are required to identify this parameter locally.
(LIV-4) is required to use the MTE to identify the standard treatment
parameters.
As demonstrated by Heckman and Vytlacil (1999, 2000, 2005)
and Heckman (2001), over the support of (X, Z), MTE can be used
to construct (under LIV-4) or bound (in the case of partial support of
P (Z)) ATE and TT. Policy-relevant treatment effects can be defined.

89

Proof: From the law of iterated expectations,


EYjX; PZ EY1 jD 1; X; PZPZ
EY0 jD 0; X; PZ1  PZ
Z 1 Z 1


y1 f y1 ; U*V jX dU*V dy1

1

PZ

1

PZ
1



y0 f y0 ; U*V jX dU*V dy0

where U*V FV (UV ). Thus



@EYjX; PZ
E Y1  Y0 jX; U*V PZ
@PZ
:
MTE

THE SCIENTIFIC MODEL OF CAUSALITY

81

LATE is a special case of this method.90 The LIV approach unifies


matching, control functions, and classical instrumental variables
under a common set of assumptions. Table 4 summarizes the alternative assumptions used in matching, control functions, and instrumental variables to identify treatment parameters identify conditional
(on X or X, Z).
4.4.4. Directed Acyclic Graphs and the Method of g-Computation
Directed acyclic graphs (DAG) (Pearl 2000) or the g-computation
algorithm (Robins 1989) have recently been advocated as mechanisms
for causal discovery. These methods improve on the method of
matching by making explicit some of the sources of the unobservables
generating the outcomes and postulating their relationships to observables. My discussion is more brief and considers only one population-level causal effect. It is based on Freedman (2001).
Figure 2, patterned after Freedman (2001), shows the essence
of the method. An unobserved confounder A is a determinant of
outcome F and variable B.91 We observe (B, C, F). Unobservables
are denoted by U. Each of (B, C, F) is assumed to be a random
variable produced in part from the variable preceding it in the triangle
and from unobservables that are assumed to be mutually independent
(hence the pattern of the arrows in Figure 2). Assume for simplicity
that A, B, C, F are discrete random variables. Figure 2 describes a
recursive model where A (UA), C and UF determine F; B and UC
determine C and UB and A (UA) determine B.
We seek to determine
PrF fjset B b
free of the unmeasured cofounder A, which affects both B and F. This is
the probability of getting F when we set B b. (Set is Pearls (2000)
do operation or Haavelmos (1943) fixing of the variables.) But
there is confounding due to A. A UA affects both B and F, so there
may be no true causal B  F relationship. How can one control for A?
90

Vytlacil (2002) establishes that LATE is a semiparametric version of a


control function estimator.
91
The symbols used in this subsection are not the same as those used in
the previous sections of this paper.

No

LIV

Yes

Marginal
Average?
(Given X, Z)

No

No

No

No (Yes in
standard
case)

Conventional, No
but not
required

No

Functional
Forms
Required?

(U0, U1, Uv) ?? ZjX


Pr(D 1jZ, X) is a nontrivial function of
Z for each X.

E(U0 D(U1  U0)jX, Z)


E(U0 D(U1  U0)jX) (ATE)
E(U0 D(U1  U0) 
E(U0 D(U1  U0)jX)jP(Z), X)
E(U0 D(U1  U0)  E(U0 D(U1
 U0)jX)jX) (TT)

E(U0jX, D 0, Z) and E(U1jX, D 1, Z)


can be varied independently
of 0 (X) and 1 (X), respectively
and intercepts can be identified through
limit arguments or symmetry assumptions

E(U1jX, D 1, Z) E(U1jX, Z)
E(U0jX, D 0, Z) E(U0jX, Z)

Key Identification
Condition for Means
(assuming separability)

*For propensity score matching, (X, Z) are replaced with P (X, Z) in defining parameters and conditioning sets.
**Conditions for writing the control function in terms of P (X, Z) are given in the text.

Yes

Yes

IV
Yes
(conventional)

Yes (for
Conventional,
nonparametric but not
identification) required

Control
Function**

No

Separability of
Observables and
Unobservables in
Outcome Equations?

No

Exclusion
Required?

Matching*

Method

TABLE 4
Identifying Assumptions and Implicit Economic Assumptions Underlying the Four Methods Discussed in this Paper
Conditional on X and Z

82
HECKMAN

83

THE SCIENTIFIC MODEL OF CAUSALITY

A = UA
(unobserved)

UB

UC

F
UF

We know
Pr (C = c | B = b)
Pr(F = f | C = c) =

Pr (F = f | A = a, C = c)Pr(A = a)
a
and

Pr(F = f | B = b) =

Pr(F = f | C = c)Pr(C = c | B = b)
c

FIGURE 2. DAG analysis. Adapted from Freedman (2001).

The g-computation algorithm operates by computing the following


probabilities based on observables. From the data, we can compute
Pr (C cjB b). We can also compute the left-hand side of
PrF fjC c

PrF fjA a; C c PrA a:

Hence we can identify the desired causal object using the following
calculation:
PrF fjset B b

X
c

PrF fjC c PrC cjB b:

84

HECKMAN

The ingredients on the right-hand side can be calculated from the


available data (recall that A is not observed).
This very useful result breaks down entirely if we add an
arrow like that shown in Figure 3, because in this case A also confounds C. The role of the a priori theory is to specify the arrows. No
purely empirical algorithm can find causal effects in general models, a
point emphasized by Freedman (2001). Figure 4 shows another case
where the g-computation approach breaks down in nonrecursive
simultaneous equations models. F  C and UF  UC interdependence create further problems ruled out in the DAG approach.
These examples all illustrate the provisional nature of causal
inference and the role of theory in justifying the estimators of causal
effects.

A = UA
(unobserved)

UB

UC

UF

FIGURE 3. If another arrow is added to Figure 2, the argument breaks down.


Where do arrows come from?

85

THE SCIENTIFIC MODEL OF CAUSALITY

A = UA
(unobserved)

UB

UC

UF

FIGURE 4. Nonrecursive. Argument breaks down. DAG is one estimation


scheme for one hypothetical model, not a general algorithm for
causal discovery.

5. SUMMARY AND CONCLUSIONS


This paper defines counterfactual models, causal parameters, and structural models and relates the parameters of the treatment effect literature to the parameters of structural econometrics and scientific causal
models. I distinguish counterfactuals from scientific causal models.
Counterfactuals are an ingredient of causal models. Scientific causal
models also specify a mechanism for selecting counterfactuals. I present
precise definitions of causal effects within structural models that are
inclusive of the specification of a mechanism (a formal model) by which
causal variables are externally manipulated (i.e., outcomes are selected).
Models of causality advocated in statistics are incomplete because they
do not specify the mechanisms of external variation that are central to
the definition of causality, nor do they specify the sources of randomness producing outcomes and the relationship between outcomes and

86

HECKMAN

selection mechanisms. By not determining the causes of effects, or


modeling the relationship between potential outcomes and
assignment to treatment, statistical models of causality cannot be
used to provide valid answers to the numerous counterfactual questions
required for policy analysis. They do not exploit relationships among
potential outcomes, assignment to treatment, and the variables causing
potential outcomes that can be used to devise econometric evaluation
estimators. The statistical approach does not model the choice of
treatment mechanism and its relationship with outcome equations,
whereas the scientific approach makes the choice of treatment
equation a centerpiece of identification analysis. The statistical model
does not apply to nonrecursive settings, whereas the econometric
model can be readily adapted to handle both recursive and nonrecursive cases.
Statistical treatment effects are typically proposed to answer a
more limited set of questions than are addressed by structural equation models and it is not surprising that they can do so under weaker
conditions than are required to identify structural equations. At the
same time, if treatment effects are used structurallythat is, to forecast the effect of a program on new populations or to forecast the
effects of new programsstronger assumptions are required of the
sort used in standard structural econometrics (see Heckman 2001;
Heckman and Vytlacil 2005, 2006b).
Table 5 compares scientific models with statistical causal models. Statistical causal models, in their current state, are not
fully articulated models. Crucial assumptions about sources of randomness are kept implicit. The assumptions required to project treatment
parameters to different populations are not specified. The scientific
approach has no substitute for making out-of-sample predictions
that is, for answering policy questions P2 and P3. The scientific
approach distinguishes derivation of a model as an abstract theoretical
activity from the problem of identifying models from data.

APPENDIX: THE VALUE OF STRUCTURAL EQUATIONS IN


MAKING POLICY FORECASTS
Structural equations are useful for three different purposes. First, the
derivatives of such functions or finite changes generate the

One focused treatment effect

Range of questions answered

Social/market interactions
Does not project

Ignored

Treatment of interdependence

Nonparametric

Recursive

Mechanism of intervention
for defining counterfactuals

Parametric?

Hypothetical randomization

Models of conditional counterfactuals

Projections to different populations?

Implicit
Implicit

Sources of randomness

Statistical Causal Models

Econometric Models

In principle, answers many possible questions


continued

Becoming nonparametric

Projects

Modeled in general equilibrium frameworks

Recursive or simultaneous systems

Many mechanisms of hypothetical


interventions including randomization;
mechanism is explicitly modeled

Explicit

Explicit

TABLE 5
Econometric Versus Statistical Causal Models

THE SCIENTIFIC MODEL OF CAUSALITY

87

88

HECKMAN

comparative statics ceteris paribus variations produced by scientific


theory. For example, tests of economic theory and measurements of
economic parameters (price elasticities, measurements of consumer
surplus, etc.) are often based on structural equations.
Second, structural equations can be used to forecast the effects
of policies evaluated in one population in other populations, provided
that the parameters are invariant across populations and that support
conditions are satisfied. However, a purely nonparametric structural
equation determined on one support cannot be extrapolated to other
populations with different supports.
Third, as emphasized by Marschak (1953), Marshallian causal
functions and structural equations are one ingredient required to
forecast the effect of a new policy, never previously implemented.
The problem of forecasting the effects of a policy evaluated
on one population but applied to another population can be formulated in the following way. Let Y(!) (X(!), U(!)), where
: D ! Y, D  R J , Y  R : is a structural equation determining outcome Y, and we assume that it is known only over
Supp(X(!), U(!)) X U: X(!) and U(!) are random input variables.
The mean outcome conditional on X(!) x is
EH YjX x

X x; udFH ujX x;
U

where FH(ujX) is the distribution of U in the historical data. We seek


to forecast the outcome in a target population that may have a different
support. The average outcome in the target population (T) is
ET YjX x

Z
UT

X x; udFT ujX x

where U T is the support of U in the target population. Provided the


support of (X, U) is the same in the source and the target populations,
from knowledge of FT it is possible to produce a correct value of
ET(YjX x) on the target population. Otherwise, it is possible to
evaluate this expectation only over the intersection set SuppT(X) \
SuppH(X), where SuppA(X) is the support of X in the A population.
In order to extrapolate over the whole set SuppT(X), it is necessary to
adopt some form of parametric or functional structure. Additive

THE SCIENTIFIC MODEL OF CAUSALITY

89

separability in simplifies the extrapolation problem. If is additively


separable
Y X U;
(X) applies to all populations for which we can condition on X.
However, some structure may have to be imposed to extrapolate
from SuppH(X) to SuppT(X) if (X) on T is not determined nonparametrically from H.
The problem of forecasting the effect of a new policy, never
previously experienced, is similar in character to the policy forecasting
problem just discussed. It shares many elements in common with the
problem of forecasting the demand for a new good, never previously
consumed.92 Without imposing some structure on this problem, it is
impossible to solve. The literature in structural econometrics associated with the work of the Cowles Commission adopts the following
five-step approach to this problem.
1.
2.

3.

4.

5.

Structural functions are determined (e.g., (X)).


The new policy is characterized by an invertible mapping from
observed random variables to the characteristics associated with
the policy: C q(X), where c is the set of characteristics associated
with the policy and q, q:RJ RJ, is a known invertible mapping.
X q1(C) is solved to associate characteristics that in principle
can be observed with the policy. This places the characteristics of
the new policy on the same footing as those of the old.
It is assumed that, in the historical data, Supp(q1(C)  Supp(X)).
This ensures that the support of the new characteristics mapped
into X space is contained in the support of X. If this condition is
not met, some functional structure must be used to forecast the
effects of the new policy, to extend it beyond the support of the
source population.
The forecast effect of the policy on Y is Y(C) (q1(C)).

92

Quandt and Baumol (1966), Lancaster (1971), Gorman (1980),


McFadden (1974), and Domencich and McFadden (1975) consider the problem
of forecasting the demand for a new good. Marschak (1953) is the classic
reference for evaluating the effect of a new policy; see Heckman (2001).

90

HECKMAN

The leading example of this approach is Lancasters method


for estimating the demand for a new good (Lancaster 1971). New
goods are viewed as bundles of old characteristics. McFaddens conditional logit scheme (1974) is based on a similar idea.93
Marschaks analysis of the effect of a new commodity tax is
another example. Let P(!) be the random variable denoting the price
facing consumer !. The tax changes the product price from P(!) to
P(!)(1 t), where t is the tax. With sufficient price variation so that
the assumption in step 4 is satisfied so that the support of the price
after tax, Supppost tax(P(!)(1 t))  Supppretax(P(!)), it is possible to
use reduced form demand functions fit on a pretax sample to forecast
the effect of a tax never previously put in place. Marschak uses a
linear structural equation to solve the problem of limited support.
From linearity, determination of the structural equations over a small
region determines it everywhere.
Marshallian or structural causal functions are an essential ingredient in constructing such forecasts because they explicitly model the
relationship between U and X. The treatment effect approach does not
explicitly model this relationship so that treatment parameters cannot be
extrapolated in this fashion, unless the dependence of potential outcomes on U and X is specified, and the required support conditions
are satisfied. The Rubin (1978)Holland (1986) model does not specify
the required relationships.

REFERENCES
Aakvik, A., J. J. Heckman, and E. J. Vytlacil. 1999. Training Effects on
Employment When the Training Effects are Heterogeneous: An Application

93
McFaddens stochastic specification is different from Lancasters specification. See Heckman and Snyder (1997) for a comparison of these two
approaches. Lancaster assumes that the U (!) are the same for each consumer
in all choice settings. (They are preference parameters in his setting.) McFadden
allows for U (!) to be different for the same consumer across different choice
settings but assumes that the U (!) in each choice setting are draws from a
common distribution that can be determined from the demand for old goods.

THE SCIENTIFIC MODEL OF CAUSALITY

91

to Norwegian Vocational Rehabilitation Programs. University of Bergen


Working Paper 0599.
. 2005. Estimating Treatment Effects for Discrete Outcomes When
Responses to Treatment Vary: An Application to Norwegian Vocational
Rehabilitation Programs. Journal of Econometrics 125(12):1551.
Abadie, A. 2003. Semiparametric Differences-in-Differences Estimators.
Department of Economics, Harvard University, Unpublished manuscript.
Abadie, A., J. D. Angrist, and G. Imbens. 2002. Instrumental Variables
Estimates of the Effect of Subsidized Training on the Quantiles of Trainee
Earnings. Econometrica 70(1):91117.
Abbring, J. H., and G. J. Van Den Berg. 2003. The Nonparametric Identification
of Treatment Effects in Duration Models. Econometrica 71(5):1491517.
Andrews, D. W., and M. M. Schafgans. 1998. Semiparametric Estimation of the
Intercept of a Sample Selection Model. Review of Economic Studies
65(3):497517.
Angrist, J. D. 1991. The Draft Lottery and Voluntary Enlistment in the Vietnam
Era. Journal of the American Statistical Association 86(415):58495.
Angrist, J. D., G. W. Imbens, and D. Rubin. 1996. Identification of Causal
Effects Using Instrumental Variables. Journal of the American Statistical
Association 91:44455.
Bjorklund, A., and R. Moffitt. 1987. The Estimation of Wage Gains and
Welfare Gains in Self-selection. Review of Economics and Statistics
69(1):4249.
Boadway, R. W., and N. Bruce. 1984. Welfare Economics. New York: Blackwell
Publishers.
Brock, W. A., and S. N. Durlauf 2001. Interactions-based models. Pp. 346368
in Handbook of Econometrics, Vol. 5, edited by J. J. Heckman and E. Leamer.
New York: North-Holland.
Cameron, S. V., and J. J. Heckman. 1998. Life Cycle Schooling and Dynamic
Selection Bias: Models and Evidence for Five Cohorts of American Males.
Journal of Political Economy 106(2):262333.
Campbell, D. T., and J. C. Stanley. 1963. Experimental and Quasi-experimental
Designs for Research. Chicago: Rand McNally.
Carneiro, P. 2002. Heterogeneity in the Returns to Schooling: Implications for
Policy Evaluation. Ph. D. dissertation, University of Chicago.
Carneiro, P., K. Hansen, and J. J. Heckman. 2001. Removing the Veil of
Ignorance in Assessing the Distributional Impacts of Social Policies.
Swedish Economic Policy Review 8(2):273301.
. 2003. Estimating Distributions of Treatment Effects with an
Application to the Returns to Schooling and Measurement of the Effects of
Uncertainty on College Choice. 2001 Lawrence R. Klein Lecture.
International Economic Review 44(2):361422.
Carneiro, P., J. J. Heckman, and E. J. Vytlacil. 2005. Understanding What
Instrumental Variables Estimate: Estimating Marginal and Average Returns
to Education. Department of Economics, University of Chicago.
Unpublished manuscript.

92

HECKMAN

Chen, S. 1999. Distribution-free Estimation of the Random Coefficient Dummy


Endogenous Variable Model. Journal of Econometrics 91(1):17199.
Cox, D. 1958. Planning of Experiments. New York: Wiley.
. 1992. Causality: Some Statistical Aspects. Journal of the Royal
Statistical Society, Series A, 155:291301.
Cox, D., and N. Wermuth. 1996. Multivariate Dependencies: Models, Analysis
and Interpretation. New York: Chapman and Hall.
Cunha, F., J. Heckman, and S. Navarro. 2005a. Counterfactual Analysis of
Inequality and Social Mobility. In Income Inequality, edited by M. Gretzky.
Palo Alto: Stanford University Press. Forthcoming.
. 2005b. Separating Heterogeneity from Uncertainty in Modeling
Schooling Choices. Oxford Economic Papers 57(2):191261.
Dawid, A. 2000. Causal Inference Without Counterfactuals. Journal of the
American Statistical Association 95(450):40724.
Domencich, T., and D. L. McFadden. 1975. Urban Travel Demand: A Behavioral
Analysis. Amsterdam: North-Holland.
Fisher, R. A. 1966. The Design of Experiments. New York: Hafner.
Florens, J.-P., and J. J. Heckman. 2003. Causality and Econometrics.
Department of Economics, University of Chicago. Unpublished working paper.
Foster, J. E., and A. K. Sen. 1998. On Economic Inequality. New York: Oxford
University Press.
Freedman, D. 2001. On Specifying Graphical Models for Causation and the
Identification Problem. Department of Statistics, University of California at
Berkeley. Unpublished manuscript.
Gill, R. D., and J. M. Robins. 2001. Causal Inference for Complex Longitudinal
Data: The Continuous Case. Annals of Statistics 29(6):17851811.
Gorman, W. M. 1980. A Possible Procedure for Analysing Quality Differentials
in the Egg Market. Review of Economic Studies 47(5):84356.
Haavelmo, T. 1943. The Statistical Implications of a System of Simultaneous
Equations. Econometrica 11(1):112.
. 1944. The Probability Approach in Econometrics. Econometrica
12(suppl.):iiivi; 1115.
Hahn, J. 1998. On the Role of the Propensity Score in Efficient Semiparametric
Estimation of Average Treatment Effects. Econometrica 66(2):31531.
Harsanyi, J. C. 1955. Cardinal Welfare, Individualistic Ethics and Interpersonal
Comparisons of Utility. Journal of Political Economy 63(4):30921.
. 1975. Can the Maximin Principle Serve as a Basis for Morality? A
Critique of John Rawlss Theory. American Political Science Review
69(2):594606.
Heckman, J. J. 1976. Simultaneous Equation Models with Both Continuous
and Discrete Endogenous Variables with and Without Structural Shift in
the Equations. Pp. 23572 in Studies in Nonlinear Estimation, edited by
S. Goldfeld and R. Quandt. Cambridge, MA: Ballinger.
. 1978. Dummy Endogenous Variables in a Simultaneous Equation
System. Econometrica 46(4):93159.

THE SCIENTIFIC MODEL OF CAUSALITY

93

. 1980. Sample Selection Bias as a Specification Error with an


Application to the Estimation of Labor Supply Functions. Pp. 20648 in
Female Labor Supply: Theory and Estimation, edited by J. P. Smith. Princeton,
NJ: Princeton University Press.
. 1990. Varieties of Selection Bias. American Economic Review 80(2),
31318.
. 1992. Randomization and Social Policy Evaluation. Pp. 20130 in
Evaluating Welfare and Training Programs, edited by C. Manski and I.
Garfinkel. Cambridge, MA: Harvard University Press.
. 1997. Instrumental Variables: A Study of Implicit Behavioral
Assumptions Used in Making Program Evaluations. Journal of Human
Resources 32(3):44162.
. 2000. Causal Parameters and Policy Analysis in Economics: A
Twentieth Century Retrospective. Quarterly Journal of Economics
115(1):4597.
. 2001. Micro Data, Heterogeneity, and the Evaluation of Public Policy:
Nobel Lecture. Journal of Political Economy 109(4):673748.
. 2006. Evaluating Economic Policy. Princeton, NJ: Princeton University
Press.
Heckman, J. J., and B. E. Honore. 1990. The Empirical Content of the Roy
Model. Econometrica 58(5):112149.
Heckman, J. J., and V. J. Hotz. 1989. Choosing Among Alternative
Nonexperimental Methods for Estimating the Impact of Social Programs:
The Case of Manpower Training. Journal of the American Statistical
Association 84(408):86274.
Heckman, J. J., H. Ichimura, J. Smith, and P. E. Todd. 1998. Characterizing
Selection Bias Using Experimental Data. Econometrica 66(5):101798.
Heckman, J. J., H. Ichimura, and P. E. Todd. 1997. Matching as an
Econometric Evaluation Estimator: Evidence from Evaluating a Job
Training Programme. Review of Economic Studies 64(4):60554.
. 1998. Matching as an Econometric Evaluation Estimator. Review of
Economic Studies 65(223):26194.
Heckman, J. J., R. J. LaLonde, and J. A. Smith. 1999. The Economics and
Econometrics of Active Labor Market Programs. Pp. 18652097 in
Handbook of Labor Economics, Vol. 3A, edited by O. Ashenfelter and
D. Card. New York: North-Holland.
Heckman, J. J., and T. E. MaCurdy. 1985. A Simultaneous Equations Linear
Probability Model. Canadian Journal of Economics 18(1):2837.
Heckman, J. J., and S. Navarro. 2004. Using Matching, Instrumental Variables,
and Control Functions to Estimate Economic Choice Models. Review of
Economics and Statistics 86(1):3057.
. 2006. Dynamic Discrete Choice and Dynamic Treatment Effects.
Journal of Econometrics. Forthcoming.
Heckman, J. J., and R. Robb. 1985. Alternative Methods for Evaluating the
Impact of Interventions. Pp. 156245 in Longitudinal Analysis of Labor

94

HECKMAN

Market Data, Vol. 10, edited by J. Heckman and B. Singer. New York:
Cambridge University Press.
. 1986. Alternative Methods for Solving the Problem of Selection Bias in
Evaluating the Impact of Treatments on Outcomes. Pp. 63107 in Drawing
Inferences from Self-Selected Samples, edited by H. Wainer. New York:
Springer-Verlag.
Heckman, J. J., and J. A. Smith. 1998. Evaluating the Welfare State. Pp. 241
318 in Econometrics and Economic Theory in the Twentieth Century: The
Ragnar Frisch Centennial Symposium, edited by S. Strom. New York:
Cambridge University Press.
Heckman, J. J., J. Smith, and N. Clements. 1997. Making the Most Out of
Programme Evaluations and Social Experiments: Accounting for
Heterogeneity in Programme Impacts. Review of Economic Studies
64(221):487536.
Heckman, J. J., and J. M. Snyder Jr. 1997. Linear Probability Models of the
Demand for Attributes with an Empirical Application to Estimating the
Preferences of Legislators (Special issue). RAND Journal of Economics
28:S142.
Heckman, J. J., J. L. Tobias, and E. J. Vytlacil. 2001. Four Parameters of
Interest in the Evaluation of Social Programs. Southern Economic Journal
68(2):21023.
. 2003. Simple Estimators for Treatment Parameters in a Latent Variable
Framework. Review of Economics and Statistics 85(3):74854.
Heckman, J. J., and E. J. Vytlacil. 1999. Local Instrumental Variables and
Latent Variable Models for Identifying and Bounding Treatment Effects.
Proceedings of the National Academy of Sciences 96:473034.
. 2000. The Relationship Between Treatment Parameters Within a Latent
Variable Framework. Economics Letters 66(1):3339.
. 2001. Local Instrumental Variables. Pp. 146 in Nonlinear Statistical
Modeling: Proceedings of the Thirteenth International Symposium in Economic
Theory and Econometrics: Essays in Honor of Takeshi Amemiya, edited by
C. Hsiao, K. Morimue, and J. L. Powell. New York: Cambridge University
Press.
. 2005. Structural Equations, Treatment Effects and Econometric Policy
Evaluation. Econometrica 73(3):669738.
. 2006a. Econometric Evaluation of Social Programs, Part I: Causal
Models, Structural Models and Econometric Policy Evaluation. In
J. Heckman and E. Leamer (Eds.), Handbook of Econometrics, Volume 6.
Amsterdam: Elsevier, forthcoming.
. 2006b. Econometric Evaluation of Social Programs, Part II: Using
Economic Choice Theory and the Marginal Treatment Effect to Organize
Alternative Econometric Estimators. In J. Heckman and E. Leamer (Eds.),
Handbook of Econometrics, Volume 6. Amsterdam: Elsevier, forthcoming.
Holland, P. W. 1986. Statistics and Causal Inference. Journal of the American
Statistical Association 81(396):94560.

THE SCIENTIFIC MODEL OF CAUSALITY

95

. 1988. Causal Inference, Path Analysis, and Recursive Structural


Equation Models. Pp. 44984 in Sociological Methodology, Vol. 18, edited
by C. Clogg and G. Arminger. Washington, DC: American Sociological
Association.
Hurwicz, L. 1962. On the Structural Form of Interdependent Systems. Pp. 232
39 in Logic, Methodology and Philosophy of Science, edited by E. Nagel,
P. Suppes, and A. Tarski. Stanford, CA: Stanford University Press.
Imbens, G. W., and J. D. Angrist. 1994. Identification and Estimation of Local
Average Treatment Effects. Econometrica 62(2):46775.
Katz, D., A. Gutek, R. Kahn, and E. Barton. 1975. Bureaucratic Encounters: A
Pilot Study in the Evaluation of Government Services. Ann Arbor: Survey
Research Center, Institute for Social Research, University of Michigan.
Knight, F. 1921. Risk, Uncertainty and Profit. New York: Houghton Mifflin.
Lancaster, K. J. 1971. Consumer Demand: A New Approach. New York:
Columbia University Press.
Leamer, E. E. 1985. Vector Autoregressions for Causal Inference? CarnegieRochester Conference Series on Public Policy 22:255303.
Lechner, M. 2004. Sequential Matching Estimation of Dynamic Causal
Models. Technical Report 2004, IZA Institute for the Study of Labor
Discussion Paper.
Lewis, H. G. 1974. Comments on Selectivity Biases in Wage Comparisons.
Journal of Political Economy 82(6):114555.
Lucas, R. E., and T. J. Sargent. 1981. Rational Expectations and Econometric
Practice. Minneapolis: University of Minnesota Press.
Marschak, J. 1953. Economic Measurements for Policy and Prediction. Pp.
126 in Studies in Econometric Method, edited by W. Hood and T. Koopmans.
New York: Wiley.
Marshall, A. 1890. Principles of Economics. New York: Macmillan.
Matzkin, R. 2003. Nonparametric Estimation of Nonadditive Random
Functions. Econometrica 71(5):133975.
. 2004. Unobserved Instruments. Department of Economics,
Northwestern University, Evanston, IL. Unpublished manuscript.
. 2006. Nonparametric Identification. In Handbook of Econometrics,
Vol. 6, edited by J. Heckman and E. Leamer. Amsterdam: Elsevier.
McFadden, D. 1974. Conditional Logit Analysis of Qualitative Choice
Behavior. In Frontiers in Econometrics, edited by P. Zarembka. New York:
Academic Press.
. 1981. Econometric Models of Probabilistic Choice. In Structural
Analysis of Discrete Data with Econometric Applications, edited by
C. Manski and D. McFadden. Cambridge, MA: MIT Press.
Mill, J. S. 1848. Principles of Political Economy with Some of Their Applications to
Social Philosophy. London: J. W. Parker.
Moulin, H. 1983. The Strategy of Social Choice. New York: North-Holland.
Neyman, J. 1923. Statistical Problems in Agricultural Experiments. Journal of
the Royal Statistical Society Series B (suppl.) (2):10780.

96

HECKMAN

Pearl, J. 2000. Causality. Cambridge, England: Cambridge University Press.


Powell, J. L. 1994. Estimation of Semiparametric Models. Pp. 2443521 in
Handbook of Econometrics, Vol. 4, edited by R. Engle and D. McFadden.
Amsterdam: Elsevier.
Quandt, R. E. 1958. The Estimation of the Parameters of a Linear Regression
System Obeying Two Separate Regimes. Journal of the American Statistical
Association 53(284):87380.
. 1972. A New Approach to Estimating Switching Regressions. Journal
of the American Statistical Association 67(338):30610.
. 1974. A Comparison of Methods for Testing Nonnested Hypotheses.
Review of Economics and Statistics 56(1):9299.
Quandt, R. E., and W. Baumol. 1966. The Demand for Abstract Transport
Modes: Theory and Measurement. Journal of Regional Science 6:1326.
Rawls, J. 1971. A Theory of Justice. Cambridge, MA: Belknap.
Robins, J. M. 1989. The Analysis of Randomized and Non-randomized AIDS
Treatment Trials Using a New Approach to Causal Inference in Longitudinal
Studies. Pp. 11359 in Health Services Research Methodology: A Focus on
AIDS, edited by L. Sechrest, H. Freeman, and A. Mulley. Rockville, MD:
U.S. Department of Health and Human Services, National Center for Health
Services Research and Health Care Technology Assessment.
. (1997). Causal Inference from Complex Longitudinal Data. Pp. 69117
in Latent Variable Modeling and Applications to Causality. Lecture Notes in
Statistics, edited by M. Berkane. New York: Springer-Verlag.
Rosenbaum, P. R. 1987. Model-Based Direct Adjustment. Journal of the
American Statistical Association 82(398):38794.
. 1995. Observational Studies. New York: Springer-Verlag.
Rosenbaum, P. R., and D. B. Rubin. 1983. The Central Role of the
Propensity Score in Observational Studies for Causal Effects. Biometrika
70(1):4155.
Roy, A. 1951. Some Thoughts on the Distribution of Earnings. Oxford
Economic Papers 3(2):13546.
Rubin, D. B. 1978. Bayesian Inference for Causal Effects: The Role of
Randomization. Annals of Statistics 6(1):3458.
. 1986. Statistics and Casual Inference: Comment: Which Ifs Have
Casual Answers. Journal of the American Statistical Association
81(396):96162.
Rubin, D. B., and N. Thomas. 1992. Characterizing the Effect of Matching
Using Linear Propensity Score Methods with Normal Distributions.
Biometrika 79(4):797809.
Ruud, P. A. 2000. An Introduction to Classical Econometric Theory. New York:
Oxford University Press.
Sen, A. K. 1999. The Possibility of Social Choice. American Economic Review
89(3):34978.
Sims, C. A. 1977. Exogeneity and Casual Orderings in Macroeconomic Models.
Pp. 2343 in New Methods in Business Cycle Research. Minneapolis, MN:
Federal Reserve Bank of Minneapolis.

THE SCIENTIFIC MODEL OF CAUSALITY

97

Tamer, E. 2003. Incomplete Simultaneous Discrete Response Model with


Multiple Equilibria. Review of Economic Studies 70(1):14765.
Thurstone, L. 1930. The Fundamentals of Statistics. New York: Macmillan.
Tukey, J. 1986. Comments on Alternative Methods for Solving the Problem
of Selection Bias in Evaluating the Impact of Treatments on Outcomes.
Pp. 10810 in Drawing Inferences from Self-Selected Samples, edited by
H. Wainer. New York: Springer-Verlag.
Vickrey, W. 1945. Measuring Marginal Utility by Reactions to Risk.
Econometrica 13(4):31933.
. 1960. Utility, Strategy, and Social Decision Rules. Quarterly Journal of
Economics 74(4):50735.
Vijverberg, W. P. M. 1993. Measuring the Unidentified Parameter of the
Extended Roy Model of Selectivity. Journal of Econometrics 57(13):6989.
Vytlacil, E. J. 2002. Independence, Monotonicity, and Latent Index Models: An
Equivalence Result. Econometrica 70(1):33141.
Wainer, H. (Ed.) 1986. Drawing Inferences from Self-Selected Samples.
New York: Springer-Verlag (Reprinted in 2000, Mahwah, NJ: Lawrence
Erlbaum Associates).

Anda mungkin juga menyukai